Sample records for observed velocity field

  1. 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.

  2. Dense Velocity Field of Turkey

    NASA Astrophysics Data System (ADS)

    Ozener, H.; Aktug, B.; Dogru, A.; Tasci, L.

    2017-12-01

    While the GNSS-based crustal deformation studies in Turkey date back to early 1990s, a homogenous velocity field utilizing all the available data is still missing. Regional studies employing different site distributions, observation plans, processing software and methodology not only create reference frame variations but also heterogeneous stochastic models. While the reference frame effect between different velocity fields could easily be removed by estimating a set of rotations, the homogenization of the stochastic models of the individual velocity fields requires a more detailed analysis. Using a rigorous Variance Component Estimation (VCE) methodology, we estimated the variance factors for each of the contributing velocity fields and combined them into a single homogenous velocity field covering whole Turkey. Results show that variance factors between velocity fields including the survey mode and continuous observations can vary a few orders of magnitude. In this study, we present the most complete velocity field in Turkey rigorously combined from 20 individual velocity 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 velocity fields. The homogenously combined new velocity 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.

  3. The 6dFGS Peculiar Velocity Field

    NASA Astrophysics Data System (ADS)

    Springob, Chris M.; Magoulas, C.; Colless, M.; Mould, J.; Erdogdu, P.; Jones, D. H.; Lucey, J.; Campbell, L.; Merson, A.; Jarrett, T.

    2012-01-01

    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 velocities, making it the largest peculiar velocity 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 velocity 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 velocity field from the 2MASS redshift survey, and compare our observed 6dFGS velocity field to the predicted field. We discuss the resulting agreement between the observed and predicted fields, and the implications for measurements of the bias parameter and bulk flow.

  4. Preflare magnetic and velocity fields

    NASA Technical Reports Server (NTRS)

    Hagyard, M. J.; Gaizauskas, V.; Chapman, G. A.; Deloach, A. C.; Gary, G. A.; Jones, H. P.; Karpen, J. T.; Martres, M.-J.; Porter, J. G.; Schmeider, B.

    1986-01-01

    A characterization is given of the preflare magnetic field, using theoretical models of force free fields together with observed field structure to determine the general morphology. Direct observational evidence for sheared magnetic fields is presented. The role of this magnetic shear in the flare process is considered within the context of a MHD model that describes the buildup of magnetic energy, and the concept of a critical value of shear is explored. The related subject of electric currents in the preflare state is discussed next, with emphasis on new insights provided by direct calculations of the vertical electric current density from vector magnetograph data and on the role of these currents in producing preflare brightenings. Results from investigations concerning velocity fields in flaring active regions, describing observations and analyses of preflare ejecta, sheared velocities, and vortical motions near flaring sites are given. This is followed by a critical review of prevalent concepts concerning the association of flux emergence with flares

  5. Observation and analysis of abrupt changes in the interplanetary plasma velocity and magnetic field.

    NASA Technical Reports Server (NTRS)

    Martin, R. N.; Belcher, J. W.; Lazarus, A. J.

    1973-01-01

    This paper presents a limited study of the physical nature of abrupt changes in the interplanetary plasma velocity and magnetic field based on 19 day's data from the Pioneer 6 spacecraft. The period was chosen to include a high-velocity solar wind stream and low-velocity wind. Abrupt events were accepted for study if the sum of the energy density in the magnetic field and velocity changes was above a specified minimum. A statistical analysis of the events in the high-velocity 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).

  6. Fabry-Perot Observations of Comet Hale-Bopp H_2O(+) Velocity Fields

    NASA Astrophysics Data System (ADS)

    Roesler, F. L.; Klinglesmith, D. A., III; Scherb, F.; Mierkiewicz, E. J.; Oliversen, R. J.

    1997-07-01

    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 observations consisted of ``data cubes,'' i.e., a sequence of images of the 6158 Angstroms emission doublet at velocity 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 velocities and accelerations on the plane of the sky.

  7. Field observations of seismic velocity changes caused by shaking-induced damage and healing due to mesoscopic nonlinearity

    NASA Astrophysics Data System (ADS)

    Gassenmeier, M.; Sens-Schönfelder, C.; Eulenfeld, T.; Bartsch, M.; Victor, P.; Tilmann, F.; Korn, M.

    2016-03-01

    To investigate temporal seismic velocity 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 velocity changes are measured with Coda Wave Interferometry. At station PATCX, we observe seasonal changes in seismic velocity caused by thermal stress as well as transient velocity reductions in the frequency range of 4-6 Hz. Sudden velocity 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 velocity variations based on continuous observations of the local ground acceleration. The model assumes that not only the shaking of large earthquakes causes velocity drops, but any small vibrations continuously induce minor velocity 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 velocity measurements, which is one day. In our model, the amplitude of the velocity 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 velocity variation in remarkable detail. Additionally, a linear trend is observed 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 velocity 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

  8. 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.

  9. On Animating 2D Velocity Fields

    NASA Technical Reports Server (NTRS)

    Kao, David; Pang, Alex; Yan, Jerry (Technical Monitor)

    2001-01-01

    A velocity field, even one that represents a steady state flow, implies a dynamical system. Animated velocity fields is an important tool in understanding such complex phenomena. This paper looks at a number of techniques that animate velocity fields and propose two new alternatives. These are texture advection and streamline cycling. The common theme among these techniques is the use of advection on some texture to generate a realistic animation of the velocity field. Texture synthesis and selection for these methods are presented. Strengths and weaknesses of the techniques are also discussed in conjunctions with several examples.

  10. On Animating 2D Velocity Fields

    NASA Technical Reports Server (NTRS)

    Kao, David; Pang, Alex

    2000-01-01

    A velocity field. even one that represents a steady state flow implies a dynamical system. Animated velocity fields is an important tool in understanding such complex phenomena. This paper looks at a number of techniques that animate velocity fields and propose two new alternatives, These are texture advection and streamline cycling. The common theme among these techniques is the use of advection on some texture to generate a realistic animation of the velocity field. Texture synthesis and selection for these methods are presented. Strengths and weaknesses of the techniques are also discussed in conjunction with several examples.

  11. POLARIZED LINE FORMATION IN NON-MONOTONIC VELOCITY FIELDS

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

    Sampoorna, M.; Nagendra, K. N., E-mail: sampoorna@iiap.res.in, E-mail: knn@iiap.res.in

    2016-12-10

    For a correct interpretation of the observed spectro-polarimetric data from astrophysical objects such as the Sun, it is necessary to solve the polarized line transfer problems taking into account a realistic temperature structure, the dynamical state of the atmosphere, a realistic scattering mechanism (namely, the partial frequency redistribution—PRD), and the magnetic fields. In a recent paper, we studied the effects of monotonic vertical velocity fields on linearly polarized line profiles formed in isothermal atmospheres with and without magnetic fields. However, in general the velocity fields that prevail in dynamical atmospheres of astrophysical objects are non-monotonic. Stellar atmospheres with shocks, multi-componentmore » supernova atmospheres, and various kinds of wave motions in solar and stellar atmospheres are examples of non-monotonic velocity fields. Here we present studies on the effect of non-relativistic non-monotonic vertical velocity fields on the linearly polarized line profiles formed in semi-empirical atmospheres. We consider a two-level atom model and PRD scattering mechanism. We solve the polarized transfer equation in the comoving frame (CMF) of the fluid using a polarized accelerated lambda iteration method that has been appropriately modified for the problem at hand. We present numerical tests to validate the CMF method and also discuss the accuracy and numerical instabilities associated with it.« less

  12. The statistical properties of sea ice velocity fields

    NASA Astrophysics Data System (ADS)

    Agarwal, S.; Wettlaufer, J. S.

    2016-12-01

    Thorndike and Colony (1982) showed that more than 70% of the variance of the ice motion can be explained by the geostrophic winds. This conclusion was reached by analyzing only 2 years of data. Due to the importance of ice motion in Arctic climate we ask how persistent is such a prediction. In so doing, we study and develop a stochastic model for the Arctic sea ice velocity fields based on the observed sea ice velocity fields from satellites and buoys for the period 1978 - 2012. Having previously found that the Arctic Sea Equivalent Ice Extent (EIE) has a white noise structure on annual to bi-annual time scales (Agarwal et. al. 2012), we assess the connection to ice motion. We divide the Arctic into dynamic and thermodynamic components, with focus on the dynamic part i.e. the velocity fields of sea ice driven by the geostrophic winds over the Arctic. We show (1) the stationarity of the spatial correlation structure of the velocity fields, and (2) the robustness of white noise structure present in the velocity fields on annual to bi-annual time scales, which combine to explain the white noise characteristics of the EIE on these time scales. S. Agarwal, W. Moon and J.S. Wettlaufer, Trends, noise and reentrant long-term persistence in Arctic sea ice, Proc. R. Soc. A, 468, 2416 (2012). A.S. Thorndike and R. Colony, Sea ice motion in response to geostrophic winds, J. Geophys. Res. 87, 5845 (1982).

  13. Excess velocity of magnetic domain walls close to the depinning field

    NASA Astrophysics Data System (ADS)

    Caballero, Nirvana B.; Fernández Aguirre, Iván; Albornoz, Lucas J.; Kolton, Alejandro B.; Rojas-Sánchez, Juan Carlos; Collin, Sophie; George, Jean Marie; Diaz Pardo, Rebeca; Jeudy, Vincent; Bustingorry, Sebastian; Curiale, Javier

    2017-12-01

    Magnetic field driven domain wall velocities in [Co/Ni] based multilayers thin films have been measured using polar magneto-optic Kerr effect microscopy. The low field results are shown to be consistent with the universal creep regime of domain wall motion, characterized by a stretched exponential growth of the velocity with the inverse of the applied field. Approaching the depinning field from below results in an unexpected excess velocity with respect to the creep law. We analyze these results using scaling theory to show that this speeding up of domain wall motion can be interpreted as due to the increase of the size of the deterministic relaxation close to the depinning transition. We propose a phenomenological model to accurately fit the observed excess velocity and to obtain characteristic values for the depinning field Hd, the depinning temperature Td, and the characteristic velocity scale v0 for each sample.

  14. The velocity field of the barred spiral galaxy NGC 1300 revisited.

    NASA Astrophysics Data System (ADS)

    Lindblad, P. A. B.; Kristen, H.; Joersaeter, S.; Hoegbom, J.

    1997-01-01

    The re-reduction, described in Joersaeter & van Moorsel (1995AJ....110.2037J), of NGC 1300 VLA HI observations, originally obtained by M. England, motivates a new analysis of the velocity field and rotation curve. Fitting tilted ring models to the HI velocity data, we find the new values for the orientation parameters of NGC 1300 to be PA_lon_=267+/-2deg and i=35+/-5deg. Subsequently, the HI rotation curve is extracted, and a residual velocity map constructed. The HI velocity residuals in the bar region are found to be consistent with elliptical motion aligned with the bar major axis. Further out the residual velocities correlate with the position of the HI spiral arms. We use 16 optical long slit emission line spectra, covering mainly the nuclear, bar, and inner arm region, to resolve the inner part of the velocity field. Three new spectra are presented in this investigation, and the remaining 13 are found in the literature. The optical velocities reveal a sharply rising rotation curve in the inner R<10", not seen in the HI data due to beam-smearing. The optical velocity field is weighted together with the HI velocities to produce a combined velocity field. This velocity field is interpreted using hydrodynamical models in a subsequent paper by Lindblad & Kristen (1996A&A...313..733L).

  15. Far-Field and Middle-Field Vertical Velocities Associated with Megathrust Earthquakes

    NASA Astrophysics Data System (ADS)

    Fleitout, L.; Trubienko, O.; Klein, E.; Vigny, C.; Garaud, J.; Shestakov, N.; Satirapod, C.; Simons, W. J.

    2013-12-01

    The recent megathrust earthquakes (Sumatra, Chili and Japan) have induced far-field postseismic subsidence with velocities from a few mm/yr to more than 1cm/yr at distances from 500 to 1500km from the earthquake epicentre, for several years following the earthquake. This subsidence is observed in Argentina, China, Korea, far-East Russia and in Malaysia and Thailand as reported by Satirapod et al. ( ASR, 2013). In the middle-field a very pronounced uplift is localized on the flank of the volcanic arc facing the trench. This is observed both over Honshu, in Chile and on the South-West coast of Sumatra. In Japan, the deformations prior to Tohoku earthquake are well measured by the GSI GPS network: While the East coast was slightly subsiding, the West coast was raising. A 3D finite element code (Zebulon-Zset) is used to understand the deformations through the seismic cycle in the areas surrounding the last three large subduction earthquakes. The meshes designed for each region feature a broad spherical shell portion with a viscoelastic asthenosphere. They are refined close to the subduction zones. Using these finite element models, we find that the pattern of the predicted far-field vertical postseismic displacements depends upon the thicknesses of the elastic plate and of the low viscosity asthenosphere. A low viscosity asthenosphere at shallow depth, just below the lithosphere is required to explain the subsidence at distances from 500 to 1500km. A thick (for example 600km) asthenosphere with a uniform viscosity predicts subsidence too far away from the trench. Slip on the subduction interface is unable tot induce the observed far-field subsidence. However, a combination of relaxation in a low viscosity wedge and slip or relaxation on the bottom part of the subduction interface is necessary to explain the observed postseismic uplift in the middle-field (volcanic arc area). The creep laws of the various zones used to explain the postseismic data can be injected in

  16. Electric field measurements during the Condor critical velocity experiment

    NASA Technical Reports Server (NTRS)

    Kelley, M. C.; Pfaff, R. F.; Haerendel, G.

    1986-01-01

    The instrumentation of the Condor critical velocity Ba experiment (Wescott et al., 1986) for the measurements of the energetic particles and the electric field associated with a Ba explosion is described. The Ba explosion created a complex electric field pulse detected in situ by a single-axis double electric-field probe on a separate spacecraft. The measurements provide evidence of several important links in the critical-velocity chain, and are consistent with two hypotheses. The first hypothesis involves the creation of large polarization electric field due to charge separation; the second hypothesis implies a polarization of the beam by currents flowing across it. The chain of physical processes inferred from the observations is in agreement with most theories for the Alfven process.

  17. 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.

  18. Inferring Lower Boundary Driving Conditions Using Vector Magnetic Field Observations

    NASA Technical Reports Server (NTRS)

    Schuck, Peter W.; Linton, Mark; Leake, James; MacNeice, Peter; Allred, Joel

    2012-01-01

    Low-beta coronal MHD simulations of realistic CME events require the detailed specification of the magnetic fields, velocities, 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 velocities such as the Differential Affine Velocity Estimator for Vector Magnetograms and the Poloidal/Toroidal Decomposition. Nominally, these velocities are consistent with the evolution of the radial magnetic field. To evolve the tangential magnetic field radial gradients must be specified. In addition to estimating the photospheric vector magnetic and velocity 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 observed magnetic fields and estimated velocities at some height within the simulation. Even if this goal is achieved, many unanswered questions remain. How can the photospheric magnetic fields and velocities be propagated to the low corona through the transition region? At what cadence must we observe the photosphere to realistically simulate the corona? How do we model the magnetic fields and plasma velocities 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?

  19. The velocity field of clusters of galaxies within 100 megaparsecs. II - Northern clusters

    NASA Technical Reports Server (NTRS)

    Mould, J. R.; Akeson, R. L.; Bothun, G. D.; Han, M.; Huchra, J. P.; Roth, J.; Schommer, R. A.

    1993-01-01

    Distances and peculiar velocities for galaxies in eight clusters and groups have been determined by means of the near-infrared Tully-Fisher relation. With the possible exception of a group halfway between us and the Hercules Cluster, we observe peculiar velocities of the same order as the measuring errors of about 400 km/s. The present sample is drawn from the northern Galactic hemisphere and delineates a quiet region in the Hubble flow. This contrasts with the large-scale flows seen in the Hydra-Centaurus and Perseus-Pisces regions. We compare the observed peculiar velocities with predictions based upon the gravity field inferred from the IRAS redshift survey. The differences between the observed and predicted peculiar motions are generally small, except near dense structures, where the observed motions exceed the predictions by significant amounts. Kinematic models of the velocity field are also compared with the data. We cannot distinguish between parameterized models with a great attractor or models with a bulk flow.

  20. Rotating field mass and velocity analyzer

    NASA Technical Reports Server (NTRS)

    Smith, Steven Joel (Inventor); Chutjian, Ara (Inventor)

    1998-01-01

    A rotating field mass and velocity analyzer having a cell with four walls, time dependent RF potentials that are applied to each wall, and a detector. The time dependent RF potentials create an RF field in the cell which effectively rotates within the cell. An ion beam is accelerated into the cell and the rotating RF field disperses the incident ion beam according to the mass-to-charge (m/e) ratio and velocity distribution present in the ion beam. The ions of the beam either collide with the ion detector or deflect away from the ion detector, depending on the m/e, RF amplitude, and RF frequency. The detector counts the incident ions to determine the m/e and velocity distribution in the ion beam.

  1. 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.

  2. 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.

  3. Neutron stars velocities and magnetic fields

    NASA Astrophysics Data System (ADS)

    Paret, Daryel Manreza; Martinez, A. Perez; Ayala, Alejandro.; Piccinelli, G.; Sanchez, A.

    2018-01-01

    We study a model that explain neutron stars velocities due to the anisotropic emission of neutrinos. Strong magnetic fields present in neutron stars are the source of the anisotropy in the system. To compute the velocity of the neutron star we model its core as composed by strange quark matter and analice the properties of a magnetized quark gas at finite temperature and density. Specifically we have obtained the electron polarization and the specific heat of magnetized fermions as a functions of the temperature, chemical potential and magnetic field which allow us to study the velocity of the neutron star as a function of these parameters.

  4. Mass-loss rates, ionization fractions, shock velocities, and magnetic fields of stellar jets

    NASA Technical Reports Server (NTRS)

    Hartigan, Patrick; Morse, Jon A.; Raymond, John

    1994-01-01

    In this paper we calculate emission-line ratios from a series of planar radiative shock models that cover a wide range of shock velocities, preshock densities, and magnetic fields. The models cover the initial conditions relevant to stellar jets, and we show how to estimate the ionization fractions and shock velocities in jets directly from observations of the strong emission lines in these flows. The ionization fractions in the HH 34, HH 47, and HH 111 jets are approximately 2%, considerably smaller than previous estimates, and the shock velocities are approximately 30 km/s. For each jet the ionization fractions were found from five different line ratios, and the estimates agree to within a factor of approximately 2. The scatter in the estimates of the shock velocities is also small (+/- 4 km/s). The low ionization fractions of stellar jets imply that the observed electron densities are much lower than the total densities, so the mass-loss rates in these flows are correspondingly higher (approximately greater than 2 x 10(exp -7) solar mass/yr). The mass-loss rates in jets are a significant fraction (1%-10%) of the disk accretion rates onto young stellar objects that drive the outflows. The momentum and energy supplied by the visible portion of a typical stellar jet are sufficient to drive a weak molecular outflow. Magnetic fields in stellar jets are difficult to measure because the line ratios from a radiative shock with a magnetic field resemble those of a lower velocity shock without a field. The observed line fluxes can in principle indicate the strength of the field if the geometry of the shocks in the jet is well known.

  5. Present-Day 3D Velocity Field of Eastern North America Based on Continuous GPS Observations

    NASA Astrophysics Data System (ADS)

    Goudarzi, Mohammad Ali; Cocard, Marc; Santerre, Rock

    2016-07-01

    The Saint Lawrence River valley in eastern Canada was studied using observations 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 velocity 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 observations. 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 velocities are radially outward from the centers of maximum uplift and are inward to the centers of maximum subsidence with the typical velocity of 1-1.6 ± 0.4 mm/year that is in agreement with the ICE-5G model to the first order.

  6. Measuring the velocity field from type Ia supernovae in an LSST-like sky survey

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

    Odderskov, Io; Hannestad, Steen, E-mail: isho07@phys.au.dk, E-mail: sth@phys.au.dk

    2017-01-01

    In a few years, the Large Synoptic Survey Telescope will vastly increase the number of type Ia supernovae observed in the local universe. This will allow for a precise mapping of the velocity field and, since the source of peculiar velocities is variations in the density field, cosmological parameters related to the matter distribution can subsequently be extracted from the velocity power spectrum. One way to quantify this is through the angular power spectrum of radial peculiar velocities on spheres at different redshifts. We investigate how well this observable can be measured, despite the problems caused by areas with nomore » information. To obtain a realistic distribution of supernovae, we create mock supernova catalogs by using a semi-analytical code for galaxy formation on the merger trees extracted from N-body simulations. We measure the cosmic variance in the velocity power spectrum by repeating the procedure many times for differently located observers, and vary several aspects of the analysis, such as the observer environment, to see how this affects the measurements. Our results confirm the findings from earlier studies regarding the precision with which the angular velocity power spectrum can be determined in the near future. This level of precision has been found to imply, that the angular velocity power spectrum from type Ia supernovae is competitive in its potential to measure parameters such as σ{sub 8}. This makes the peculiar velocity power spectrum from type Ia supernovae a promising new observable, which deserves further attention.« less

  7. 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.

  8. Response of Velocity Anisotropy of Shale Under Isotropic and Anisotropic Stress Fields

    NASA Astrophysics Data System (ADS)

    Li, Xiaying; Lei, Xinglin; Li, Qi

    2018-03-01

    We investigated the responses of P-wave velocity and associated anisotropy in terms of Thomsen's parameters to isotropic and anisotropic stress fields on Longmaxi shales cored along different directions. An array of piezoelectric ceramic transducers allows us to measure P-wave velocities along numerous different propagation directions. Anisotropic parameters, including the P-wave velocity α along a symmetry axis, Thomsen's parameters ɛ and δ, and the orientation of the symmetry axis, could then be extracted by fitting Thomsen's weak anisotropy model to the experimental data. The results indicate that Longmaxi shale displays weakly intrinsic velocity anisotropy with Thomsen's parameters ɛ and δ being approximately 0.05 and 0.15, respectively. The isotropic stress field has only a slight effect on velocity and associated anisotropy in terms of Thomsen's parameters. In contrast, both the magnitude and orientation of the anisotropic stress field with respect to the shale fabric are important in controlling the evolution of velocity and associated anisotropy in a changing stress field. For shale with bedding-parallel loading, velocity anisotropy is enhanced because velocities with smaller angles relative to the maximum stress increase significantly during the entire loading process, whereas those with larger angles increase slightly before the yield stress and afterwards decrease with the increasing differential stress. For shale with bedding-normal loading, anisotropy reversal is observed, and the anisotropy is progressively modified by the applied differential stress. Before reaching the yield stress, velocities with smaller angles relative to the maximum stress increase more significantly and even exceed the level of those with larger angles. After reaching the yield stress, velocities with larger angles decrease more significantly. Microstructural features such as the closure and generation of microcracks can explain the modification of the velocity anisotropy

  9. 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

  10. The velocity field of a coronal mass ejection - The event of September 1, 1980

    NASA Technical Reports Server (NTRS)

    Low, B. C.; Hundhausen, A. J.

    1987-01-01

    The velocity field of a mass ejection that was observed by the coronagraph of the SMM satellite over the northwest limb of the sun at about 0600 UT on September 1, 1980 is studied in detail. A descriptive account of the event is given, concentrating on qualitative features of the mass motion and suggesting a possible origin of the unusual two-loop structure. The velocity field is analyzed quantitatively, and the implications of the results for the mass ejection theory are considered. It is concluded that a self-similar description of the velocity field is a gross oversimplification and that although some evidence of wave propagation can be found, the bright features in the mass ejection are plasma structures moving with frozen-in magnetic fields, rather than waves propagating through plasmas and magnetic fields.

  11. 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.

  12. Estimating cosmic velocity fields from density fields and tidal tensors

    NASA Astrophysics Data System (ADS)

    Kitaura, Francisco-Shu; Angulo, Raul E.; Hoffman, Yehuda; Gottlöber, Stefan

    2012-10-01

    In this work we investigate the non-linear and non-local relation between cosmological density and peculiar velocity fields. Our goal is to provide an algorithm for the reconstruction of the non-linear velocity field from the fully non-linear density. We find that including the gravitational tidal field tensor using second-order Lagrangian perturbation theory based upon an estimate of the linear component of the non-linear density field significantly improves the estimate of the cosmic flow in comparison to linear theory not only in the low density, but also and more dramatically in the high-density regions. In particular we test two estimates of the linear component: the lognormal model and the iterative Lagrangian linearization. The present approach relies on a rigorous higher order Lagrangian perturbation theory analysis which incorporates a non-local relation. It does not require additional fitting from simulations being in this sense parameter free, it is independent of statistical-geometrical optimization and it is straightforward and efficient to compute. The method is demonstrated to yield an unbiased estimator of the velocity field on scales ≳5 h-1 Mpc with closely Gaussian distributed errors. Moreover, the statistics of the divergence of the peculiar velocity field is extremely well recovered showing a good agreement with the true one from N-body simulations. The typical errors of about 10 km s-1 (1σ confidence intervals) are reduced by more than 80 per cent with respect to linear theory in the scale range between 5 and 10 h-1 Mpc in high-density regions (δ > 2). We also find that iterative Lagrangian linearization is significantly superior in the low-density regime with respect to the lognormal model.

  13. Magnetic field and radial velocities of the star Chi Draconis A

    NASA Astrophysics Data System (ADS)

    Lee, Byeong-Cheol; Gadelshin, D.; Han, Inwoo; Kang, Dong-Il; Kim, Kang-Min; Valyavin, G.; Galazutdinov, G.; Jeong, Gwanghui; Beskrovnaya, N.; Burlakova, T.; Grauzhanina, A.; Ikhsanov, N. R.; Kholtygin, A. F.; Valeev, A.; Bychkov, V.; Park, Myeong-Gu

    2018-01-01

    We present high-resolution spectropolarimetric observations of the spectroscopic binary χ Dra. Spectral lines in the spectrum of the main component χ Dra A show variable Zeeman displacement, which confirms earlier suggestions about the presence of a weak magnetic field on the surface of this star. Within about 2 yr of time base of our observations, the longitudinal component BL of the magnetic field exhibits variation from -11.5 ± 2.5 to +11.1 ± 2.1 G with a period of about 23 d. Considering the rotational velocity of χ Dra A in the literature and that newly measured in this work, this variability may be explained by the stellar rotation under the assumption that the magnetic field is globally stable. Our new measurements of the radial velocities (RV) in high-resolution I-spectra of χ Dra A refined the orbital parameters and reveal persistent deviations of RVs from the orbital curve. We suspect that these deviations may be due to the influence of local magnetically generated spots, pulsations, or a Jupiter-size planet orbiting the system.

  14. Field-effect transistor having a superlattice channel and high carrier velocities at high applied fields

    DOEpatents

    Chaffin, R.J.; Dawson, L.R.; Fritz, I.J.; Osbourn, G.C.; Zipperian, T.E.

    1984-04-19

    In a field-effect transistor comprising a semiconductor having therein a source, a drain, a channel and a gate in operational relationship, there is provided an improvement wherein said semiconductor is a superlattice comprising alternating quantum well and barrier layers, the quantum well layers comprising a first direct gap semiconductor material which in bulk form has a certain bandgap and a curve of electron velocity versus applied electric field which has a maximum electron velocity at a certain electric field, the barrier layers comprising a second semiconductor material having a bandgap wider than that of said first semiconductor material, wherein the layer thicknesses of said quantum well and barrier layers are sufficiently thin that the alternating layers constitute a superlattice having a curve of electron velocity versus applied electric field which has a maximum electron velocity at a certain electric field, and wherein the thicknesses of said quantum well layers are selected to provide a superlattice curve of electron velocity versus applied electric field whereby, at applied electric fields higher than that at which the maximum electron velocity occurs in said first material when in bulk form, the electron velocities are higher in said superlattice than they are in said first semiconductor material in bulk form.

  15. The height variation of supergranular velocity fields determined from simultaneous OSO 8 satellite and ground-based observations

    NASA Technical Reports Server (NTRS)

    November, L. J.; Toomre, J.; Gebbie, K. B.; Simon, G. W.

    1979-01-01

    Results are reported for simultaneous satellite and ground-based observations of supergranular velocities 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. Observations of the steady Doppler velocities seen toward the limb in the middle chromosphere and the photosphere are compared; the observed 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 observations imply that supergranular velocities should be evident in the transition region and that strong horizontal shear layers in supergranulation should produce turbulence and internal gravity waves.

  16. Normalized velocity profiles of field-measured turbidity currents

    USGS Publications Warehouse

    Xu, Jingping

    2010-01-01

    Multiple turbidity currents were recorded in two submarine canyons with maximum speed as high as 280 cm/s. For each individual turbidity current measured at a fixed station, its depth-averaged velocity typically decreased over time while its thickness increased. Some turbidity currents gained in speed as they traveled downcanyon, suggesting a possible self-accelerating process. The measured velocity profiles, first in this high resolution, allowed normalizations with various schemes. Empirical functions, obtained from laboratory experiments whose spatial and time scales are two to three orders of magnitude smaller, were found to represent the field data fairly well. The best similarity collapse of the velocity profiles was achieved when the streamwise velocity and the elevation were normalized respectively by the depth-averaged velocity and the turbidity current thickness. This normalization scheme can be generalized to an empirical function Y = exp(–αXβ) for the jet region above the velocity maximum. Confirming theoretical arguments and laboratory results of other studies, the field turbidity currents are Froude-supercritical.

  17. The velocity field of growing ear cartilage.

    PubMed Central

    Cox, R W; Peacock, M A

    1978-01-01

    The velocity vector field of the growing rabbit ear cartilage has been investigated between 12 and 299 days. Empirical curves have been computed for path lines and for velocities between 12 and 87 days. The tissue movement has been found to behave as an irrotational flow of material. Stream lines and velocity equipotential lines have been calculated and provide akinematic description of the changes during growth. The importance of a knowledge of the velocity vector in physical descriptions of growth and morphological differentiation at the tissue and cellular levels is emphasized. PMID:689993

  18. Eye Velocity Gain Fields in MSTd During Optokinetic Stimulation

    PubMed Central

    Brostek, Lukas; Büttner, Ulrich; Mustari, Michael J.; Glasauer, Stefan

    2015-01-01

    Lesion studies argue for an involvement of cortical area dorsal medial superior temporal area (MSTd) in the control of optokinetic response (OKR) eye movements to planar visual stimulation. Neural recordings during OKR suggested that MSTd neurons directly encode stimulus velocity. On the other hand, studies using radial visual flow together with voluntary smooth pursuit eye movements showed that visual motion responses were modulated by eye movement-related signals. Here, we investigated neural responses in MSTd during continuous optokinetic stimulation using an information-theoretic approach for characterizing neural tuning with high resolution. We show that the majority of MSTd neurons exhibit gain-field-like tuning functions rather than directly encoding one variable. Neural responses showed a large diversity of tuning to combinations of retinal and extraretinal input. Eye velocity-related activity was observed prior to the actual eye movements, reflecting an efference copy. The observed tuning functions resembled those emerging in a network model trained to perform summation of 2 population-coded signals. Together, our findings support the hypothesis that MSTd implements the visuomotor transformation from retinal to head-centered stimulus velocity signals for the control of OKR. PMID:24557636

  19. Critical points of the cosmic velocity field and the uncertainties in the value of the Hubble constant

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

    Liu, Hao; Naselsky, Pavel; Mohayaee, Roya, E-mail: liuhao@nbi.dk, E-mail: roya@iap.fr, E-mail: naselsky@nbi.dk

    2016-06-01

    The existence of critical points for the peculiar velocity field is a natural feature of the correlated vector field. These points appear at the junctions of velocity domains with different orientations of their averaged velocity vectors. Since peculiar velocities are the important cause of the scatter in the Hubble expansion rate, we propose that a more precise determination of the Hubble constant can be made by restricting analysis to a subsample of observational data containing only the zones around the critical points of the peculiar velocity field, associated with voids and saddle points. On large-scales the critical points, where themore » first derivative of the gravitational potential vanishes, can easily be identified using the density field and classified by the behavior of the Hessian of the gravitational potential. We use high-resolution N-body simulations to show that these regions are stable in time and hence are excellent tracers of the initial conditions. Furthermore, we show that the variance of the Hubble flow can be substantially minimized by restricting observations to the subsample of such regions of vanishing velocity instead of aiming at increasing the statistics by averaging indiscriminately using the full data sets, as is the common approach.« less

  20. Electric-field control of magnetic domain-wall velocity in ultrathin cobalt with perpendicular magnetization.

    PubMed

    Chiba, D; Kawaguchi, M; Fukami, S; Ishiwata, N; Shimamura, K; Kobayashi, K; Ono, T

    2012-06-06

    Controlling the displacement of a magnetic domain wall is potentially useful for information processing in magnetic non-volatile memories and logic devices. A magnetic domain wall can be moved by applying an external magnetic field and/or electric current, and its velocity depends on their magnitudes. Here we show that the applying an electric field can change the velocity of a magnetic domain wall significantly. A field-effect device, consisting of a top-gate electrode, a dielectric insulator layer, and a wire-shaped ferromagnetic Co/Pt thin layer with perpendicular anisotropy, was used to observe it in a finite magnetic field. We found that the application of the electric fields in the range of ± 2-3 MV cm(-1) can change the magnetic domain wall velocity in its creep regime (10(6)-10(3) m s(-1)) by more than an order of magnitude. This significant change is due to electrical modulation of the energy barrier for the magnetic domain wall motion.

  1. An algorithm to estimate unsteady and quasi-steady pressure fields from velocity field measurements.

    PubMed

    Dabiri, John O; Bose, Sanjeeb; Gemmell, Brad J; Colin, Sean P; Costello, John H

    2014-02-01

    We describe and characterize a method for estimating the pressure field corresponding to velocity field measurements such as those obtained by using particle image velocimetry. The pressure gradient is estimated from a time series of velocity fields for unsteady calculations or from a single velocity field for quasi-steady calculations. The corresponding pressure field is determined based on median polling of several integration paths through the pressure gradient 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 velocity 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.

  2. Kinematic Clues to OB Field Star Origins: Radial Velocities, Runaways, and Binaries

    NASA Astrophysics Data System (ADS)

    Januszewski, Helen; Castro, Norberto; Oey, Sally; Becker, Juliette; Kratter, Kaitlin M.; Mateo, Mario; Simón-Díaz, Sergio; Bjorkman, Jon E.; Bjorkman, Karen; Sigut, Aaron; Smullen, Rachel; M2FS Team

    2018-01-01

    Field OB stars are a crucial probe of star formation in extreme conditions. Properties of massive stars formed in relative isolation can distinguish between competing star formation theories, while the statistics of runaway stars allow an indirect test of the densest conditions in clusters. To address these questions, we have obtained multi-epoch, spectroscopic observations for a spatially complete sample of 48 OB field stars in the SMC Wing with the IMACS and M2FS multi-object spectrographs at the Magellan Telescopes. The observations span 3-6 epochs per star, with sampling frequency ranging from one day to about one year. From these spectra, we have calculated the radial velocities (RVs) and, in particular, the systemic velocities for binaries. Thus, we present the intrinsic RV distribution largely uncontaminated by binary motions. We estimate the runaway frequency, corresponding to the high velocity stars in our sample, and we also constrain the binary frequency. The binary frequency and fitted orbital parameters also place important constraints on star formation theories, as these properties drive the process of runaway ejection in clusters, and we discuss these properties as derived from our sample. This unique kinematic analysis of a high mass field star population thus provides a new look at the processes governing formation and interaction of stars in environments at extreme densities, from isolation to dense clusters.

  3. A Unified Geodetic Vertical Velocity Field (UGVVF), Version 1.0

    NASA Astrophysics Data System (ADS)

    Schmalzle, G.; Wdowinski, S.

    2014-12-01

    Tectonic motion, volcanic inflation or deflation, as well as oil, gas and water pumping can induce vertical motion. In southern California these signals are inter-mingled. In tectonics, properly identifying regions that are contaminated by other signals can be important when estimating fault slip rates. Until recently vertical deformation rates determined by high precision Global Positioning Systems (GPS) had large uncertainties compared to horizontal components and were rarely used to constrain tectonic models of fault motion. However, many continuously occupied GPS stations have been operating for ten or more years, often delivering uncertainties of ~1 mm/yr or less, providing better constraints for tectonic modeling. Various processing centers produced GPS time series and estimated vertical velocity fields, each with their own set of processing techniques and assumptions. We compare vertical velocity solutions estimated by seven data processing groups as well as two combined solutions (Figure 1). These groups include: Central Washington University (CWU) and New Mexico Institute of Technology (NMT), and their combined solution provided by the Plate Boundary Observatory (PBO) through the UNAVCO website. Also compared are the Jet Propulsion Laboratory (JPL) and Scripps Orbit and Permanent Array Center (SOPAC) and their combined solution provided as part of the NASA MEaSUREs project. Smaller velocity fields included are from Amos et al., 2014, processed at the Nevada Geodetic Laboratory, Shen et al., 2011, processed by UCLA and called the Crustal Motion Map 4.0 (CMM4) dataset, and a new velocity field provided by the University of Miami (UM). Our analysis includes estimating and correcting for systematic vertical velocity and uncertainty differences between groups. Our final product is a unified velocity field that contains the median values of the adjusted velocity fields and their uncertainties. This product will be periodically updated when new velocity fields

  4. Using observations of slipping velocities to test the hypothesis that reconnection heats the active region corona

    NASA Astrophysics Data System (ADS)

    Yang, Kai; Longcope, Dana; Guo, Yang; Ding, Mingde

    2017-08-01

    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 observed rate. In the absence of coronal reconnection, field line footpoints move at the same velocity as the plasma they find themselves in. The rate of coronal reconnection is therefore related to any discrepancy observed between footpoint motion and that of the local plasma — so-called slipping motion. We propose a novel method to measure this velocity discrepancy by combining a sequence of non-linear force-free field extrapolations with maps of photospheric velocity. 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 observed slipping velocity 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 observations of the same AR made by SDO. We also compare differential emission measure inferred from those observations to that of the modeled corona.

  5. Velocity field calculation for non-orthogonal numerical grids

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

    Flach, G. P.

    2015-03-01

    Computational grids containing cell faces that do not align with an orthogonal (e.g. Cartesian, cylindrical) coordinate system are routinely encountered in porous-medium numerical simulations. Such grids are referred to in this study as non-orthogonal grids because some cell faces are not orthogonal to a coordinate system plane (e.g. xy, yz or xz plane in Cartesian coordinates). Non-orthogonal grids are routinely encountered at the Savannah River Site in porous-medium flow simulations for Performance Assessments and groundwater flow modeling. Examples include grid lines that conform to the sloping roof of a waste tank or disposal unit in a 2D Performance Assessment simulation,more » and grid surfaces that conform to undulating stratigraphic surfaces in a 3D groundwater flow model. Particle tracking is routinely performed after a porous-medium numerical flow simulation to better understand the dynamics of the flow field and/or as an approximate indication of the trajectory and timing of advective solute transport. Particle tracks are computed by integrating the velocity field from cell to cell starting from designated seed (starting) positions. An accurate velocity field is required to attain accurate particle tracks. However, many numerical simulation codes report only the volumetric flowrate (e.g. PORFLOW) and/or flux (flowrate divided by area) crossing cell faces. For an orthogonal grid, the normal flux at a cell face is a component of the Darcy velocity vector in the coordinate system, and the pore velocity for particle tracking is attained by dividing by water content. For a non-orthogonal grid, the flux normal to a cell face that lies outside a coordinate plane is not a true component of velocity with respect to the coordinate system. Nonetheless, normal fluxes are often taken as Darcy velocity components, either naively or with accepted approximation. To enable accurate particle tracking or otherwise present an accurate depiction of the velocity field for a

  6. The Seismo-Generated Electric Field Probed by the Ionospheric Ion Velocity

    NASA Astrophysics Data System (ADS)

    (Tiger) Liu, Jann-Yenq

    2017-04-01

    The ion density, ion temperature, and the ion velocity probed by IPEI (ionospheric Plasma and Electrodynamics Instrument) onboard ROCSAT (i.e. FORMOSAT-1), and the global ionospheric map (GIM) of the total electron content (TEC) derived from measurements of ground-based GPS receivers are employed to study seismo-ionospheric precursors (SIPs) of the 31 March 2002 M6.8 Earthquake in Taiwan. The GIM TEC and ROCSAT/IPEI ion density significantly decrease specifically over the epicenter area 1-5 days before the earthquake, which suggests that the associated SIPs have observed. The ROCSAT/IPEI ion temperature reveals no significant changes before and after the earthquake, while the latitude-time-TEC plots extracted from the GIMs along the Taiwan longitude illustrate that the equatorial ionization anomaly significantly weakens and moves equatorward, which indicates that the daily dynamo electric field has been disturbed and cancelled by possible seismo-generated electric field on 2 days before (29 March) the earthquake. Here, for the first time a vector parameter of ion velocity is employed to study SIPs. It is found that ROCSAT/IPEI ion velocity becomes significantly downward, which confirms that a westward electric field of about 0.91mV/m generated during the earthquake preparation period being essential 1-5 days before the earthquake. Liu, J. Y., and C. K. Chao (2016), An observing system simulation experiment for FORMOSAT-5/AIP detecting seismo-ionospheric precursors, Terrestrial Atmospheric and Oceanic Sciences, DOI: 10.3319/TAO.2016.07.18.01(EOF5).

  7. Potential, velocity, and density fields from sparse and noisy redshift-distance samples - Method

    NASA Technical Reports Server (NTRS)

    Dekel, Avishai; Bertschinger, Edmund; Faber, Sandra M.

    1990-01-01

    A method for recovering the three-dimensional potential, velocity, and density fields from large-scale redshift-distance samples is described. Galaxies are taken as tracers of the velocity field, not of the mass. The density field and the initial conditions are calculated using an iterative procedure that applies the no-vorticity assumption at an initial time and uses the Zel'dovich approximation to relate initial and final positions of particles on a grid. The method is tested using a cosmological N-body simulation 'observed' at the positions of real galaxies in a redshift-distance sample, taking into account their distance measurement errors. Malmquist bias and other systematic and statistical errors are extensively explored using both analytical techniques and Monte Carlo simulations.

  8. Assimilation of drifters' trajectories in velocity fields from coastal radar and model via the Lagrangian assimilation algorithm LAVA.

    NASA Astrophysics Data System (ADS)

    Berta, Maristella; Bellomo, Lucio; Griffa, Annalisa; Gatimu Magaldi, Marcello; Marmain, Julien; Molcard, Anne; Taillandier, Vincent

    2013-04-01

    The Lagrangian assimilation algorithm LAVA (LAgrangian Variational Analysis) is customized for coastal areas in the framework of the TOSCA (Tracking Oil Spills & Coastal Awareness network) Project, to improve the response to maritime accidents in the Mediterranean Sea. LAVA assimilates drifters' trajectories in the velocity fields which may come from either coastal radars or numerical models. In the present study, LAVA is applied to the coastal area in front of Toulon (France). Surface currents are available from a WERA radar network (2km spatial resolution, every 20 minutes) and from the GLAZUR model (1/64° spatial resolution, every hour). The cluster of drifters considered is constituted by 7 buoys, transmitting every 15 minutes for a period of 5 days. Three assimilation cases are considered: i) correction of the radar velocity field, ii) correction of the model velocity field and iii) reconstruction of the velocity field from drifters only. It is found that drifters' trajectories compare well with the ones obtained by the radar and the correction to radar velocity field is therefore minimal. Contrarily, observed and numerical trajectories separate rapidly and the correction to the model velocity field is substantial. For the reconstruction from drifters only, the velocity fields obtained are similar to the radar ones, but limited to the neighbor of the drifter paths.

  9. Velocity shear Kelvin-Helmholtz instability with inhomogeneous DC electric field in the magnetosphere of Saturn

    NASA Astrophysics Data System (ADS)

    Kandpal, Praveen; Kaur, Rajbir; Pandey, R. S.

    2018-01-01

    In this paper parallel flow velocity 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), velocity shear scale length (Ai), temperature anisotropy (T⊥ /T||), electric field (E), ratio of electron to ion temperature (Te /Ti), density gradient (εnρi) and angle of propagation (θ) on the dimensionless growth rate of K-H waves in the inner magnetosphere of Saturn has been observed 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 velocity shear, temperature anisotropy and magnitude of the electric field are observed 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 gradient have been observed 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.

  10. Recovering the full velocity and density fields from large-scale redshift-distance samples

    NASA Technical Reports Server (NTRS)

    Bertschinger, Edmund; Dekel, Avishai

    1989-01-01

    A new method for extracting the large-scale three-dimensional velocity and mass density fields from measurements of the radial peculiar velocities is presented. Galaxies are assumed to trace the velocity field rather than the mass. The key assumption made is that the Lagrangian velocity field has negligible vorticity, as might be expected from perturbations that grew by gravitational instability. By applying the method to cosmological N-body simulations, it is demonstrated that it accurately reconstructs the velocity field. This technique promises a direct determination of the mass density field and the initial conditions for the formation of large-scale structure from galaxy peculiar velocity surveys.

  11. 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

  12. Experimental study of the free surface velocity field in an asymmetrical confluence

    NASA Astrophysics Data System (ADS)

    Creelle, Stephan; Mignot, Emmanuel; Schindfessel, Laurent; De Mulder, Tom

    2017-04-01

    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 velocity 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 velocity difference between the most contracted section and the separation zone is pronounced. With the help of surface particle tracking velocimetry the velocity 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 velocity profile) that interact with the stagnation zone and each other, causing a shear layer between the two bulk flows. Close observation of the velocity 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 observed 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 gradient, velocity difference and width of the shear layer are calculated. Analysis of these data shows that the shear layer between the incoming flows

  13. 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.

  14. 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

  15. Impacts of distinct observations during the 2009 Prince William Sound field experiment: A data assimilation study

    NASA Astrophysics Data System (ADS)

    Li, Zhijin; Chao, Yi; Farrara, John D.; McWilliams, James C.

    2013-07-01

    A set of data assimilation experiments, known as Observing System Experiments (OSEs) are performed to assess the relative impacts of different types of observations acquired during the 2009 Prince William Sound Field Experiment. The observations assimilated consist primarily of two types: High Frequency (HF) radar surface velocities and vertical profiles of temperature/salinity (T/S) measured by ships, moorings, an Autonomous Underwater Vehicle and a glider. The impact of all the observations, HF radar surface velocities, and T/S profiles is assessed. Without data assimilation, a frequently occurring cyclonic eddy in the central Sound is overly persistent and intense. The assimilation of the HF radar velocities effectively reduces these biases and improves the representation of the velocities as well as the T/S fields in the Sound. The assimilation of the T/S profiles improves the large scale representation of the temperature/salinity and also the velocity field in the central Sound. The combination of the HF radar surface velocities and sparse T/S profiles results in an observing system capable of representing the circulation in the Sound reliably and thus producing analyses and forecasts with useful skill.

  16. Phase-field simulations of velocity selection in rapidly solidified binary alloys

    NASA Astrophysics Data System (ADS)

    Fan, Jun; Greenwood, Michael; Haataja, Mikko; Provatas, Nikolas

    2006-09-01

    Time-dependent simulations of two-dimensional isothermal Ni-Cu dendrites are simulated using a phase-field model solved with a finite-difference adaptive mesh refinement technique. Dendrite tip velocity selection is examined and found to exhibit a transition between two markedly different regimes as undercooling is increased. At low undercooling, the dendrite tip growth rate is consistent with the kinetics of the classical Stefan problem, where the interface is assume to be in local equilibrium. At high undercooling, the growth velocity selected approaches a linear dependence on melt undercooling, consistent with the continuous growth kinetics of Aziz and with a one-dimensional steady-state phase-field asymptotic analysis of Ahmad [Phys. Rev. E 58, 3436 (1998)]. Our simulations are also consistent with other previously observed behaviors of dendritic growth as undercooling is increased. These include the transition of dendritic morphology to absolute stability and nonequilibrium solute partitioning. Our results show that phase-field models of solidification, which inherently contain a nonzero interface width, can be used to study the dynamics of complex solidification phenomena involving both equilibrium and nonequilibrium interface growth kinetics.

  17. Three-Dimensional Velocity Field De-Noising using Modal Projection

    NASA Astrophysics Data System (ADS)

    Frank, Sarah; Ameli, Siavash; Szeri, Andrew; Shadden, Shawn

    2017-11-01

    PCMRI and Doppler ultrasound are common modalities for imaging velocity 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 velocity 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 velocity gradient throughout the domain while enforcing a divergence-free condition. The measured velocity 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.

  18. Linear velocity fields in non-Gaussian models for large-scale structure

    NASA Technical Reports Server (NTRS)

    Scherrer, Robert J.

    1992-01-01

    Linear velocity fields in two types of physically motivated non-Gaussian models are examined for large-scale structure: seed models, in which the density field is a convolution of a density profile with a distribution of points, and local non-Gaussian fields, derived from a local nonlinear transformation on a Gaussian field. The distribution of a single component of the velocity is derived for seed models with randomly distributed seeds, and these results are applied to the seeded hot dark matter model and the global texture model with cold dark matter. An expression for the distribution of a single component of the velocity in arbitrary local non-Gaussian models is given, and these results are applied to such fields with chi-squared and lognormal distributions. It is shown that all seed models with randomly distributed seeds and all local non-Guassian models have single-component velocity distributions with positive kurtosis.

  19. 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).

  20. Vertical rise velocity of equatorial plasma bubbles estimated from Equatorial Atmosphere Radar (EAR) observations and HIRB model simulations

    NASA Astrophysics Data System (ADS)

    Tulasi Ram, S.; Ajith, K. K.; Yokoyama, T.; Yamamoto, M.; Niranjan, K.

    2017-06-01

    The vertical rise velocity (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 observed out of which 68 were postsunset EPBs and remaining 18 EPBs were observed around midnight hours. The vertical rise velocities of the EPBs observed around the midnight hours are significantly smaller ( 26-128 m/s) compared to those observed 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 velocity and maximum attainable altitudes of EPBs. The estimated rise velocities from EAR observations at both postsunset and midnight hours are, in general, consistent with the nonlinear evolution of EPBs from the HIRB model. The smaller vertical rise velocities (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.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 <span class="hlt">field</span> lines to latitudes away from the equator. The knowledge on bubble rise <span class="hlt">velocities</span> and their</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_1");'>1</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li class="active"><span>3</span></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_3 --> <div id="page_4" 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_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</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="61"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014ChPhB..23k4702W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014ChPhB..23k4702W"><span>Experimental studies on flow visualization and <span class="hlt">velocity</span> <span class="hlt">field</span> of compression ramp with different incoming 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>Wu, Yu; Yi, Shi-He; He, Lin; Chen, Zhi; Zhu, Yang-Zhu</p> <p>2014-11-01</p> <p>Experimental studies which focus on flow visualization and the <span class="hlt">velocity</span> <span class="hlt">field</span> of a supersonic laminar/turbulent flow over a compression ramp were carried out in a Mach 3.0 wind tunnel. Fine flow structures and <span class="hlt">velocity</span> <span class="hlt">field</span> structures were obtained via NPLS (nanoparticle-tracer planar laser scattering) and PIV (particle image velocimetry) techniques, time-averaged flow structures were researched, and spatiotemporal evolutions of transient flow structures were analyzed. The flow visualization results indicated that when the ramp angles were 25°, a typical separation occurred in the laminar flow, some typical flow structures such as shock induced by the boundary layer, separation shock, reversed flow and reattachment shock were visible clearly. While a certain extent separation occurred in turbulent flow, the separation region was much smaller. When the ramp angles were 28°, laminar flow separated further, and the separation region expanded evidently, flow structures in the separation region were complex. While a typical separation occurred in turbulent flow, reversed flow structures were significant, flow structures in the separation region were relatively simple. The experimental results of <span class="hlt">velocity</span> <span class="hlt">field</span> were corresponding to flow visualization, and the <span class="hlt">velocity</span> <span class="hlt">field</span> structures of both compression ramp flows agreed with the flow structures well. There were three layered structures in the U component <span class="hlt">velocity</span>, and the V component <span class="hlt">velocity</span> appeared like an oblique “v”. Some differences between these two compression ramp flows can be <span class="hlt">observed</span> in the <span class="hlt">velocity</span> profiles of the shear layer and the shearing intensity.</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 <span class="hlt">field</span> 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('https://ntrs.nasa.gov/search.jsp?R=20060035786&hterms=alicia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dalicia%2Bd','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20060035786&hterms=alicia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dalicia%2Bd"><span>Magnetic Cloud <span class="hlt">Field</span> Intensities and Solar Wind <span class="hlt">Velocities</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Gonzalez, Walter D.; Clau de Gonzalez, Alicia D.; Tsurutani, Bruce T.; Arballo, John K.</p> <p>1997-01-01</p> <p>For the sets of magnetic clouds studied in this work we have shown that there is a general relationship between their magnetic <span class="hlt">fields</span> strength and <span class="hlt">velocities</span>. With a clear tendency that the faster the speed of the cloud the higher the magnetic <span class="hlt">field</span>.</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> <span class="hlt">Field</span> 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 <span class="hlt">field</span> 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 <span class="hlt">field</span> 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> <span class="hlt">field</span>. 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 gradient 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 gradient 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/2015AGUFM.G13A0995T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.G13A0995T"><span>Cluster Analysis of <span class="hlt">Velocity</span> <span class="hlt">Field</span> Derived from Dense GNSS Network of Japan</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takahashi, A.; Hashimoto, M.</p> <p>2015-12-01</p> <p>Dense GNSS networks have been widely used to <span class="hlt">observe</span> crustal deformation. Simpson et al. (2012) and Savage and Simpson (2013) have conducted cluster analyses of GNSS <span class="hlt">velocity</span> <span class="hlt">field</span> in the San Francisco Bay Area and Mojave Desert, respectively. They have successfully found <span class="hlt">velocity</span> discontinuities. They also showed an advantage of cluster analysis for classifying GNSS <span class="hlt">velocity</span> <span class="hlt">field</span>. Since in western United States, strike-slip events are dominant, geometry is simple. However, the Japanese Islands are tectonically complicated due to subduction of oceanic plates. There are many types of crustal deformation such as slow slip event and large postseismic deformation. We propose a modified clustering method of GNSS <span class="hlt">velocity</span> <span class="hlt">field</span> in Japan to separate time variant and static crustal deformation. Our modification is performing cluster analysis every several months or years, then qualifying cluster member similarity. If a GNSS station moved differently from its neighboring GNSS stations, the station will not belong to in the cluster which includes its surrounding stations. With this method, time variant phenomena were distinguished. We applied our method to GNSS data of Japan from 1996 to 2015. According to the analyses, following conclusions were derived. The first is the clusters boundaries are consistent with known active faults. For examples, the Arima-Takatsuki-Hanaore fault system and the Shimane-Tottori segment proposed by Nishimura (2015) are recognized, though without using prior information. The second is improving detectability of time variable phenomena, such as a slow slip event in northern part of Hokkaido region detected by Ohzono et al. (2015). The last one is the classification of postseismic deformation caused by large earthquakes. The result suggested <span class="hlt">velocity</span> discontinuities in postseismic deformation of the Tohoku-oki earthquake. This result implies that postseismic deformation is not continuously decaying proportional to distance from its epicenter.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JCoPh.323...75R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JCoPh.323...75R"><span>Creating analytically divergence-free <span class="hlt">velocity</span> <span class="hlt">fields</span> from grid-based data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ravu, Bharath; Rudman, Murray; Metcalfe, Guy; Lester, Daniel R.; Khakhar, Devang V.</p> <p>2016-10-01</p> <p>We present a method, based on B-splines, to calculate a C2 continuous analytic vector potential from discrete 3D <span class="hlt">velocity</span> data on a regular grid. A continuous analytically divergence-free <span class="hlt">velocity</span> <span class="hlt">field</span> can then be obtained from the curl of the potential. This <span class="hlt">field</span> can be used to robustly and accurately integrate particle trajectories in incompressible flow <span class="hlt">fields</span>. Based on the method of Finn and Chacon (2005) [10] this new method ensures that the analytic <span class="hlt">velocity</span> <span class="hlt">field</span> matches the grid values almost everywhere, with errors that are two to four orders of magnitude lower than those of existing methods. We demonstrate its application to three different problems (each in a different coordinate system) and provide details of the specifics required in each case. We show how the additional accuracy of the method results in qualitatively and quantitatively superior trajectories that results in more accurate identification of Lagrangian coherent structures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhTea..56..114P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhTea..56..114P"><span>Measuring average angular <span class="hlt">velocity</span> with a smartphone magnetic <span class="hlt">field</span> sensor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pili, Unofre; Violanda, Renante</p> <p>2018-02-01</p> <p>The angular <span class="hlt">velocity</span> of a spinning object is, by standard, measured using a device called a tachometer. However, by directly using it in a classroom setting, the activity is likely to appear as less instructive and less engaging. Indeed, some alternative classroom-suitable methods for measuring angular <span class="hlt">velocity</span> have been presented. In this paper, we present a further alternative that is smartphone-based, making use of the real-time magnetic <span class="hlt">field</span> (simply called B-<span class="hlt">field</span> in what follows) data gathering capability of the B-<span class="hlt">field</span> sensor of the smartphone device as the timer for measuring average rotational period and average angular <span class="hlt">velocity</span>. The in-built B-<span class="hlt">field</span> sensor in smartphones has already found a number of uses in undergraduate experimental physics. For instance, in elementary electrodynamics, it has been used to explore the well-known Bio-Savart law and in a measurement of the permeability of air.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AdWR...94..470S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AdWR...94..470S"><span>Mass-conservative reconstruction of Galerkin <span class="hlt">velocity</span> <span class="hlt">fields</span> for transport simulations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Scudeler, C.; Putti, M.; Paniconi, C.</p> <p>2016-08-01</p> <p>Accurate calculation of mass-conservative <span class="hlt">velocity</span> <span class="hlt">fields</span> from numerical solutions of Richards' equation is central to reliable surface-subsurface flow and transport modeling, for example in long-term tracer simulations to determine catchment residence time distributions. In this study we assess the performance of a local Larson-Niklasson (LN) post-processing procedure for reconstructing mass-conservative <span class="hlt">velocities</span> from a linear (P1) Galerkin finite element solution of Richards' equation. This approach, originally proposed for a-posteriori error estimation, modifies the standard finite element <span class="hlt">velocities</span> by imposing local conservation on element patches. The resulting reconstructed flow <span class="hlt">field</span> is characterized by continuous fluxes on element edges that can be efficiently used to drive a second order finite volume advective transport model. Through a series of tests of increasing complexity that compare results from the LN scheme to those using <span class="hlt">velocity</span> <span class="hlt">fields</span> derived directly from the P1 Galerkin solution, we show that a locally mass-conservative <span class="hlt">velocity</span> <span class="hlt">field</span> is necessary to obtain accurate transport results. We also show that the accuracy of the LN reconstruction procedure is comparable to that of the inherently conservative mixed finite element approach, taken as a reference solution, but that the LN scheme has much lower computational costs. The numerical tests examine steady and unsteady, saturated and variably saturated, and homogeneous and heterogeneous cases along with initial and boundary conditions that include dry soil infiltration, alternating solute and water injection, and seepage face outflow. Typical problems that arise with <span class="hlt">velocities</span> derived from P1 Galerkin solutions include outgoing solute flux from no-flow boundaries, solute entrapment in zones of low hydraulic conductivity, and occurrences of anomalous sources and sinks. In addition to inducing significant mass balance errors, such manifestations often lead to oscillations in concentration</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22455518','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22455518"><span><span class="hlt">Observed</span> differences in upper extremity forces, muscle efforts, postures, <span class="hlt">velocities</span> and accelerations across computer activities in a <span class="hlt">field</span> study of office workers.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bruno Garza, J L; Eijckelhof, B H W; Johnson, P W; Raina, S M; Rynell, P W; Huysmans, M A; van Dieën, J H; van der Beek, A J; Blatter, B M; Dennerlein, J T</p> <p>2012-01-01</p> <p>This study, a part of the PRedicting Occupational biomechanics in OFfice workers (PROOF) study, investigated whether there are differences in <span class="hlt">field</span>-measured forces, muscle efforts, postures, <span class="hlt">velocities</span> and accelerations across computer activities. These parameters were measured continuously for 120 office workers performing their own work for two hours each. There were differences in nearly all forces, muscle efforts, postures, <span class="hlt">velocities</span> and accelerations across keyboard, mouse and idle activities. Keyboard activities showed a 50% increase in the median right trapezius muscle effort when compared to mouse activities. Median shoulder rotation changed from 25 degrees internal rotation during keyboard use to 15 degrees external rotation during mouse use. Only keyboard use was associated with median ulnar deviations greater than 5 degrees. Idle activities led to the greatest variability <span class="hlt">observed</span> in all muscle efforts and postures measured. In future studies, measurements of computer activities could be used to provide information on the physical exposures experienced during computer use. Practitioner Summary: Computer users may develop musculoskeletal disorders due to their force, muscle effort, posture and wrist <span class="hlt">velocity</span> and acceleration exposures during computer use. We report that many physical exposures are different across computer activities. This information may be used to estimate physical exposures based on patterns of computer activities over time.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUFM.T31B1819T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFM.T31B1819T"><span>Two-dimensional, average <span class="hlt">velocity</span> <span class="hlt">field</span> across the Asal Rift, Djibouti from 1997-2008 RADARSAT data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tomic, J.; Doubre, C.; Peltzer, G.</p> <p>2009-12-01</p> <p>Located at the western end of the Aden ridge, the Asal Rift is the first emerged section of the ridge propagating into Afar, a region of intense volcanic and tectonic activity. We construct a two-dimensional surface <span class="hlt">velocity</span> map of the 200x400 km2 region covering the rift using the 1997-2008 archive of InSAR data acquired from ascending and descending passes of the RADARSAT satellite. The large phase signal due to turbulent troposphere conditions over the Afar region is mostly removed from the 11-year average line of sight (LOS) <span class="hlt">velocity</span> maps, revealing a clear deformation signal across the rift. We combine the ascending and descending pass LOS <span class="hlt">velocity</span> <span class="hlt">fields</span> with the Arabia-Somalia pole of rotation adjusted to regional GPS <span class="hlt">velocities</span> (Vigny et al., 2007) to compute the <span class="hlt">fields</span> of the vertical and horizontal, GPS-parallel components of the <span class="hlt">velocity</span> over the rift. The vertical <span class="hlt">velocity</span> <span class="hlt">field</span> shows a ~40 km wide zone of doming centered over the Fieale caldera associated with shoulder uplift and subsidence of the rift inner floor. Differential movement between shoulders and floor is accommodated by creep at 6 mm/yr on Fault γ and 2.7 mm/yr on Fault E. The horizontal <span class="hlt">field</span> shows that the two shoulders open at a rate of ~15 mm/yr, while the horizontal <span class="hlt">velocity</span> decreases away from the rift to the plate motion rate of ~11 mm/yr. Part of the opening is concentrated on faults γ (5 mm/yr) and E (4 mm/yr) and about 4 mm/yr is distributed between Fault E and Fault H in the southern part of the rift. The <span class="hlt">observed</span> <span class="hlt">velocity</span> <span class="hlt">field</span> along a 60 km-long profile across the eastern part of the rift can be explained with a 2D mechanical model involving a 5-9 km-deep, vertical dyke expanding horizontally at a rate of 5 cm/yr, a 2 km-wide, 7 km-deep sill expanding vertically at 1cm/yr, and down-dip and opening of faults γ and E. Results from 3D rift models describing along-strike <span class="hlt">velocity</span> decrease away from the Goubbet Gulf and the effects of a pressurized magma chamber will be</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-<span class="hlt">field</span>, rather than far-<span class="hlt">field</span> 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://www.ncbi.nlm.nih.gov/pubmed/23231111','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23231111"><span>Sound <span class="hlt">field</span> separation with sound pressure and particle <span class="hlt">velocity</span> measurements.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fernandez-Grande, Efren; Jacobsen, Finn; Leclère, Quentin</p> <p>2012-12-01</p> <p>In conventional near-<span class="hlt">field</span> acoustic holography (NAH) it is not possible to distinguish between sound from the two sides of the array, thus, it is a requirement that all the sources are confined to only one side and radiate into a free <span class="hlt">field</span>. When this requirement cannot be fulfilled, sound <span class="hlt">field</span> separation techniques make it possible to distinguish between outgoing and incoming waves from the two sides, and thus NAH can be applied. In this paper, a separation method based on the measurement of the particle <span class="hlt">velocity</span> in two layers and another method based on the measurement of the pressure and the <span class="hlt">velocity</span> in a single layer are proposed. The two methods use an equivalent source formulation with separate transfer matrices for the outgoing and incoming waves, so that the sound from the two sides of the array can be modeled independently. A weighting scheme is proposed to account for the distance between the equivalent sources and measurement surfaces and for the difference in magnitude between pressure and <span class="hlt">velocity</span>. Experimental and numerical studies have been conducted to examine the methods. The double layer <span class="hlt">velocity</span> method seems to be more robust to noise and flanking sound than the combined pressure-<span class="hlt">velocity</span> method, although it requires an additional measurement surface. On the whole, the separation methods can be useful when the disturbance of the incoming <span class="hlt">field</span> is significant. Otherwise the direct reconstruction is more accurate and straightforward.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFMOS54A..03L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFMOS54A..03L"><span>Impacts of distinct <span class="hlt">observations</span> during the 2009 Prince William Sound <span class="hlt">field</span> experiment: A data assimilation study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Z.; Chao, Y.; Farrara, J.; McWilliams, J. C.</p> <p>2012-12-01</p> <p>A set of data assimilation experiments, known as <span class="hlt">Observing</span> System Experiments (OSEs), are performed to assess the relative impacts of different types of <span class="hlt">observations</span> acquired during the 2009 Prince William Sound <span class="hlt">Field</span> Experiment. The <span class="hlt">observations</span> assimilated consist primarily of three types: High Frequency (HF) radar surface <span class="hlt">velocities</span>, vertical profiles of temperature/salinity (T/S) measured by ships, moorings, Autonomous Underwater Vehicles and gliders, and satellite sea surface temperatures (SSTs). The impact of all the <span class="hlt">observations</span>, HF radar surface <span class="hlt">velocities</span>, and T/S profiles is assessed. Without data assimilation, a frequently occurring cyclonic eddy in the central Sound is overly persistent and intense. The assimilation of the HF radar <span class="hlt">velocities</span> effectively reduces these biases and improves the representation of the <span class="hlt">velocities</span> as well as the T/S <span class="hlt">fields</span> in the Sound. The assimilation of the T/S profiles improves the large scale representation of the temperature/salinity and also the <span class="hlt">velocity</span> <span class="hlt">field</span> in the central Sound. The combination of the HF radar surface <span class="hlt">velocities</span> and sparse T/S profiles results in an <span class="hlt">observing</span> system capable of representing the circulation in the Sound reliably and thus producing analyses and forecasts with useful skill. It is suggested that a potentially promising <span class="hlt">observing</span> network could be based on satellite SSHs and SSTs along with sparse T/S profiles, and future satellite SSHs with wide swath coverage and higher resolution may offer excellent data that will be of great use for predicting the circulation in the Sound.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ExFl...57...76K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ExFl...57...76K"><span><span class="hlt">Velocity</span> <span class="hlt">field</span> measurements on high-frequency, supersonic microactuators</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kreth, Phillip A.; Ali, Mohd Y.; Fernandez, Erik J.; Alvi, Farrukh S.</p> <p>2016-05-01</p> <p>The resonance-enhanced microjet actuator which was developed at the Advanced Aero-Propulsion Laboratory at Florida State University is a fluidic-based device that produces pulsed, supersonic microjets by utilizing a number of microscale, flow-acoustic resonance phenomena. The microactuator used in this study consists of an underexpanded source jet that flows into a cylindrical cavity with a single, 1-mm-diameter exhaust orifice through which an unsteady, supersonic jet issues at a resonant frequency of 7 kHz. The flowfields of a 1-mm underexpanded free jet and the microactuator are studied in detail using high-magnification, phase-locked flow visualizations (microschlieren) and two-component particle image velocimetry. These are the first direct measurements of the <span class="hlt">velocity</span> <span class="hlt">fields</span> produced by such actuators. Comparisons are made between the flow visualizations and the <span class="hlt">velocity</span> <span class="hlt">field</span> measurements. The results clearly show that the microactuator produces pulsed, supersonic jets with <span class="hlt">velocities</span> exceeding 400 m/s for roughly 60 % of their cycles. With high unsteady momentum output, this type of microactuator has potential in a range of ow control applications.</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 <span class="hlt">field</span> 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://www.ncbi.nlm.nih.gov/pubmed/27075772','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27075772"><span>Effect of Range and Angular <span class="hlt">Velocity</span> of Passive Movement on Somatosensory Evoked Magnetic <span class="hlt">Fields</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sugawara, Kazuhiro; Onishi, Hideaki; Yamashiro, Koya; Kojima, Sho; Miyaguchi, Shota; Kotan, Shinichi; Tsubaki, Atsuhiro; Kirimoto, Hikari; Tamaki, Hiroyuki; Shirozu, Hiroshi; Kameyama, Shigeki</p> <p>2016-09-01</p> <p>To clarify characteristics of each human somatosensory evoked <span class="hlt">field</span> (SEF) component following passive movement (PM), PM1, PM2, and PM3, using high spatiotemporal resolution 306-channel magnetoencephalography and varying PM range and angular <span class="hlt">velocity</span>. We recorded SEFs following PM under three conditions [normal range-normal <span class="hlt">velocity</span> (NN), small range-normal <span class="hlt">velocity</span> (SN), and small range-slow <span class="hlt">velocity</span> (SS)] with changing movement range and angular <span class="hlt">velocity</span> in 12 participants and calculated the amplitude, equivalent current dipole (ECD) location, and the ECD strength for each component. All components were <span class="hlt">observed</span> in six participants, whereas only PM1 and PM3 in the other six. Clear response deflections at the ipsilateral hemisphere to PM side were <span class="hlt">observed</span> in seven participants. PM1 amplitude was larger under NN and SN conditions, and mean ECD location for PM1 was at primary motor area. PM3 amplitude was larger under SN condition and mean ECD location for PM3 under SS condition was at primary somatosensory area. PM1 amplitude was dependent on the angular <span class="hlt">velocity</span> of PM, suggesting that PM1 reflects afferent input from muscle spindle, whereas PM3 amplitude was dependent on the duration. The ECD for PM3 was located in the primary somatosensory cortex, suggesting that PM3 reflects cutaneous input. We confirmed the hypothesis for locally distinct generators and characteristics of each SEF component.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.tmp.1566H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.tmp.1566H"><span>Magnetic <span class="hlt">field</span> and radial <span class="hlt">velocities</span> of the star β CrB</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Han, Inwoo; Valyavin, G.; Galazutdinov, G.; Plachinda, S.; Butkovskaya, V.; Lee, B. C.; Kim, Kang-Min; Jeong, Gwanghui; Romanyuk, I.; Burlakova, T.</p> <p>2018-06-01</p> <p>We present the results of long term high-resolution spectral and spectro-polarimetric <span class="hlt">observations</span> of the classic spectroscopic binary star β Coronae Borealis, which is also well known for its chemical peculiarity and strong magnetic <span class="hlt">field</span>. One of the main objectives of these <span class="hlt">observations</span> was to check some suggestions of the presence of a third low-mass component in the system. Analysing our own radial <span class="hlt">velocity</span> measurements of β Coronae Borealis obtained between 2004 and 2013 together with other RV and astrometric data taken from literature, we have considerably improved the orbital elements of the star. Using residual RVs obtained after subtracting the orbit from the <span class="hlt">observed</span> RV, we detected a significant periodic signal consistent with the well known 18.4868 day rotation/magnetic period of the star. The shape of the residual RVs folded with the rotation period exhibits a bimodal structure. In order to clarify the origin of the residual RV variation, we model the RV variation due to rotational modulation Zeeman patterns in the spectral lines of β Coronae Borealis spectra. The results of this analysis and the investigation of chemical inhomogeneities in the spectra of β Coronae Borealis convincingly support the "magnetic" nature of the <span class="hlt">observed</span> radial <span class="hlt">velocity</span> variations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007ExFl...42..847F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007ExFl...42..847F"><span>Simultaneous <span class="hlt">velocity</span> and concentration <span class="hlt">field</span> measurements of passive-scalar mixing in a confined rectangular jet</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Feng, Hua; Olsen, Michael G.; Hill, James C.; Fox, Rodney O.</p> <p>2007-06-01</p> <p>Simultaneous <span class="hlt">velocity</span> and concentration <span class="hlt">fields</span> in a confined liquid-phase rectangular jet with a Reynolds number based on the hydraulic diameter of 50,000 (or 10,000 based on the <span class="hlt">velocity</span> difference between streams and the jet exit dimension) and a Schmidt number of 1,250 were obtained by means of a combined particle image velocimetry (PIV) and planar laser-induced fluorescence (PLIF) system. Data were collected at the jet exit and six further downstream locations. The <span class="hlt">velocity</span> and concentration <span class="hlt">field</span> data were analyzed for flow statistics such as turbulent fluxes, turbulent viscosity and diffusivity, and turbulent Schmidt number ( Sc T ). The streamwise turbulent flux was found to be larger than the transverse turbulent flux, and the mean concentration gradient was not aligned with the turbulent flux vector. The average Sc T was found to vary both in streamwise and in cross stream directions and had a mean value around 0.8, a value consistent with the literature. Spatial correlation <span class="hlt">fields</span> of turbulent fluxes and concentration were then determined. The R u'ϕ' correlation was elliptical in shape with a major axis tilted downward with respect to the streamwise axis, whereas the R v'ϕ' correlation was an ellipse with a major axis aligned with the cross-stream direction. Negative regions of R u'ϕ' were <span class="hlt">observed</span> in the outer streams, and these negatively correlated regions decayed with downstream distance and finally disappeared altogether. The R ϕ'ϕ' correlation <span class="hlt">field</span> was found to be an ellipse with the major axis inclined at about 45° with respect to the streamwise direction. Linear stochastic estimation was used to interpret spatial correlation data and to determine conditional flow structures. It is believed that a vortex street formed near the splitter plate is responsible for the negatively correlated region <span class="hlt">observed</span> in the R u'ϕ' spatial correlations of turbulent fluxes. A positive concentration fluctuation event was <span class="hlt">observed</span> to correspond to a finger of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19860057107&hterms=Electric+current&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DElectric%2Bcurrent','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19860057107&hterms=Electric+current&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DElectric%2Bcurrent"><span>Relationship of the interplanetary electric <span class="hlt">field</span> to the high-latitude ionospheric electric <span class="hlt">field</span> and currents <span class="hlt">Observations</span> and model simulation</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Clauer, C. R.; Banks, P. M.</p> <p>1986-01-01</p> <p>The electrical coupling between the solar wind, magnetosphere, and ionosphere is studied. The coupling is analyzed using <span class="hlt">observations</span> of high-latitude ion convection measured by the Sondre Stromfjord radar in Greenland and a computer simulation. The computer simulation calculates the ionospheric electric potential distribution for a given configuration of <span class="hlt">field</span>-aligned currents and conductivity distribution. The technique for measuring F-region in <span class="hlt">velocities</span> at high time resolution over a large range of latitudes is described. Variations in the currents on ionospheric plasma convection are examined using a model of <span class="hlt">field</span>-aligned currents linking the solar wind with the dayside, high-latitude ionosphere. The data reveal that high-latitude ionospheric convection patterns, electric <span class="hlt">fields</span>, and <span class="hlt">field</span>-aligned currents are dependent on IMF orientation; it is <span class="hlt">observed</span> that the electric <span class="hlt">field</span>, which drives the F-region plasma curve, responds within about 14 minutes to IMF variations in the magnetopause. Comparisons of the simulated plasma convection with the ion <span class="hlt">velocity</span> measurements reveal good correlation between the data.</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> <span class="hlt">field</span> 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> </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_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_4 --> <div id="page_5" 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_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li class="active"><span>5</span></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</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="81"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20080047102&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=20080047102&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> <span class="hlt">Field</span> 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>Makarov, V. V.; Murphy, D. W.</p> <p>2007-01-01</p> <p>We analyze the local <span class="hlt">field</span> 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 <span class="hlt">field</span> 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 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...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19770019111','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19770019111"><span>The mean magnetic <span class="hlt">field</span> of the sun: <span class="hlt">Observations</span> at Stanford</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Scherrer, P. H.; Wilcox, J. M.; Svalgaard, L.; Duvall, T. L., Jr.; Dittmer, P. H.; Gustafson, E. K.</p> <p>1977-01-01</p> <p>A solar telescope was built at Stanford University to study the organization and evolution of large-scale solar magnetic <span class="hlt">fields</span> and <span class="hlt">velocities</span>. The <span class="hlt">observations</span> are made using a Babcock-type magnetograph which is connected to a 22.9 m vertical Littrow spectrograph. Sun-as-a-star integrated light measurements of the mean solar magnetic <span class="hlt">field</span> were made daily since May 1975. The typical mean <span class="hlt">field</span> magnitude is about 0.15 gauss with typical measurement error less than 0.05 gauss. The mean <span class="hlt">field</span> polarity pattern is essentially identical to the interplanetary magnetic <span class="hlt">field</span> sector structure (seen near the earth with a 4 day lag). The differences in the <span class="hlt">observed</span> structures can be understood in terms of a warped current sheet model.</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> <span class="hlt">field</span> 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('https://www.osti.gov/biblio/5801837-rocketborne-observations-ion-convection-electric-fields-dayside-nightside-visual-auroral-arcs','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5801837-rocketborne-observations-ion-convection-electric-fields-dayside-nightside-visual-auroral-arcs"><span>Rocketborne <span class="hlt">observations</span> of ion convection and electric <span class="hlt">fields</span> in dayside and nightside visual auroral arcs</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>Yau, A.W.; Whalen, B.A.; Creutzberg, F.</p> <p>1981-08-01</p> <p>We present ionospheric ion convection measurements in a series of four rocket payloads in and near dayside and nightside auroral arcs: one at Cape Parry (75.4/sup 0/N invariant latitude) near 1300 MLT and three at Churchill (70.0/sup 0/N invariant latitude) between 1900 and 2200 MLT. Direct measurements were made of the ionospheric ion <span class="hlt">velocity</span> distribution function, and the <span class="hlt">observed</span> ion convection <span class="hlt">velocities</span> and equivalent convective electric <span class="hlt">fields</span> were correlated with the energetic particle precipitation, the optical morphology of the aurora, and the topology of the geomagnetic <span class="hlt">field</span>. Both in the postnoon and premidnight sectors it was <span class="hlt">observed</span> that (1) equatorwardmore » of the region(s) of precipitation the ion flow was predominantly westward, with <span class="hlt">velocity</span> of about 1 km/s; (2) poleward of the region(s) the flow was predominantly westward, with <span class="hlt">velocity</span> of about 1 km/s; (2) poleward of the region(s) the flow was predominantly eastward: (3) the change in the flow direction, where <span class="hlt">observed</span>, occurred near though not exactly at the edges of the precipitation region; (4) the flow inside the precipitation region was lower; (5) the reversal of the ion flow, where <span class="hlt">observed</span>, occurred on closed magnetic <span class="hlt">field</span> lines; and (6) the convective electric <span class="hlt">field</span> typically dropped from 40 to 80 mV/m outside the precipitation region to 10 to 30 mV/m within. In the dayside Cape Perry flight, where quantitative photometric measurements were available, detailed anticorrelation between the ion convection speed and the green line emission intensity was also <span class="hlt">observed</span>.« less</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 <span class="hlt">field</span> 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> <span class="hlt">field</span> 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> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29219385','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29219385"><span>Magnetospheric Multiscale <span class="hlt">Observation</span> of Plasma <span class="hlt">Velocity</span>-Space Cascade: Hermite Representation and Theory.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>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</p> <p>2017-11-17</p> <p>Plasma turbulence is investigated using unprecedented high-resolution ion <span class="hlt">velocity</span> distribution measurements by the Magnetospheric Multiscale mission (MMS) in the Earth's magnetosheath. This novel <span class="hlt">observation</span> of a highly structured particle distribution suggests a cascadelike process in <span class="hlt">velocity</span> space. Complex <span class="hlt">velocity</span> space structure is investigated using a three-dimensional Hermite transform, revealing, for the first time in <span class="hlt">observational</span> 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 <span class="hlt">observations</span>. The combined use of state-of-the-art MMS data sets, novel implementation of a Hermite transform method, and scaling theory of the <span class="hlt">velocity</span> cascade opens new pathways to the understanding of plasma turbulence and the crucial <span class="hlt">velocity</span> space features that lead to dissipation in plasmas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..MAR.C7013J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..MAR.C7013J"><span>Experimental <span class="hlt">observations</span> of low-<span class="hlt">velocity</span> collisional systems</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, Joshua</p> <p></p> <p>Low-<span class="hlt">velocity</span> 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-<span class="hlt">velocity</span> 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 <span class="hlt">velocity</span>. For impact <span class="hlt">velocities</span> in the range from about 20-100 cm/s we <span class="hlt">observe</span> that mantling of particles has the most significant effect, reducing the coefficients of restitution.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AstBu..72..391R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AstBu..72..391R"><span>Results of magnetic <span class="hlt">field</span> measurements performed with the 6-m telescope. IV. <span class="hlt">Observations</span> in 2010</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Romanyuk, I. I.; Semenko, E. A.; Kudryavtsev, D. O.; Moiseeva, A. V.; Yakunin, I. A.</p> <p>2017-10-01</p> <p>We present the results of measurements of magnetic <span class="hlt">fields</span>, radial <span class="hlt">velocities</span> and rotation <span class="hlt">velocities</span> for 92 objects, mainly main-sequence chemically peculiar stars. <span class="hlt">Observations</span> were performed at the 6-m BTA telescope using Main Stellar Spectrograph with a Zeeman analyzer. In 2010, twelve new magnetic stars were discovered: HD 17330, HD 29762, HD 49884, HD 54824, HD 89069, HD 96003, HD 113894, HD 118054, HD 135679, HD 138633, HD 138777, BD +53.1183. The presence of a <span class="hlt">field</span> is suspected in HD 16705, HD 35379 and HD 35881. <span class="hlt">Observations</span> of standard stars without a magnetic <span class="hlt">field</span> confirm the absence of systematic errors which can introduce distortions into the measurements of longitudinal <span class="hlt">field</span>. The paper gives comments on the results of investigation of each star.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003GeoRL..30.2164N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003GeoRL..30.2164N"><span>Dynamically balanced absolute sea level of the global ocean derived from near-surface <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>Niiler, Pearn P.; Maximenko, Nikolai A.; McWilliams, James C.</p> <p>2003-11-01</p> <p>The 1992-2002 time-mean absolute sea level distribution of the global ocean is computed for the first time from <span class="hlt">observations</span> of near-surface <span class="hlt">velocity</span>. For this computation, we use the near-surface horizontal momentum balance. The <span class="hlt">velocity</span> <span class="hlt">observed</span> 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 <span class="hlt">field</span> has now been derived from <span class="hlt">observations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70016884','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70016884"><span>An exact solution of solute transport by one-dimensional random <span class="hlt">velocity</span> <span class="hlt">fields</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Cvetkovic, V.D.; Dagan, G.; Shapiro, A.M.</p> <p>1991-01-01</p> <p>The problem of one-dimensional transport of passive solute by a random steady <span class="hlt">velocity</span> <span class="hlt">field</span> is investigated. This problem is representative of solute movement in porous media, for example, in vertical flow through a horizontally stratified formation of variable porosity with a constant flux at the soil surface. Relating moments of particle travel time and displacement, exact expressions for the advection and dispersion coefficients in the Focker-Planck equation are compared with the perturbation results for large distances. The first- and second-order approximations for the dispersion coefficient are robust for a lognormal <span class="hlt">velocity</span> <span class="hlt">field</span>. The mean Lagrangian <span class="hlt">velocity</span> is the harmonic mean of the Eulerian <span class="hlt">velocity</span> for large distances. This is an artifact of one-dimensional flow where the continuity equation provides for a divergence free fluid flux, rather than a divergence free fluid <span class="hlt">velocity</span>. ?? 1991 Springer-Verlag.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1817627S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1817627S"><span>Seismological <span class="hlt">Field</span> <span class="hlt">Observation</span> of 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>Sens-Schönfelder, Christoph; Gassenmeier, Martina; Eulenfeld, Tom; Tilmann, Frederik; Korn, Michael; Niederleithinger, Ernst</p> <p>2016-04-01</p> <p>Noise based <span class="hlt">observations</span> of seismic <span class="hlt">velocity</span> changes have been made in various environments. We know of seasonal changes of <span class="hlt">velocities</span> related to ground water or temperature changes, co-seismic changes originating from shaking or stress redistribution and changes related to volcanic activity. Is is often argued that a decrease of <span class="hlt">velocity</span> is related to the opening of cracks while the closure of cracks leads to a <span class="hlt">velocity</span> increase if permanent stress changes are invoked. In contrast shaking induced changes are often related to "damage" and subsequent "healing" of the material. The co-seismic decrease and transient recovery of seismic <span class="hlt">velocities</span> can thus be explained with both - static stress changes or damage/healing processes. This results in ambiguous interpretations of the <span class="hlt">observations</span>. Here we present the analysis of one particular seismic station in northern Chile that shows very strong and clear <span class="hlt">velocity</span> changes associated with several earthquakes ranging from Mw=5.3 to Mw=8.1. The fact that we can <span class="hlt">observe</span> the response to several events of various magnitudes from different directions offers the unique possibility to discern the two possible causative processes. We test the hypothesis, that the <span class="hlt">velocity</span> changes are related to shaking rather than stress changes by developing an empirical model that is based on the local ground acceleration at the sensor site. The eight year of almost continuous <span class="hlt">observations</span> of <span class="hlt">velocity</span> changes are well modeled by a daily drop of the <span class="hlt">velocity</span> followed by an exponential recovery. Both, the amplitude of the drop as well as the recovery time are proportional to the integrated acceleration at the seismic station. Effects of consecutive days are independent and superimposed resulting in strong changes after earthquakes and constantly increasing <span class="hlt">velocities</span> during quiet days thereafter. This model describes the continuous <span class="hlt">observations</span> of the <span class="hlt">velocity</span> changes solely based on the acceleration time series without individually defined dates</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> <span class="hlt">field</span> 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/2017GeoJI.208.1088W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GeoJI.208.1088W"><span>Present-day <span class="hlt">velocity</span> <span class="hlt">field</span> and block kinematics of Tibetan Plateau from GPS measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Wei; Qiao, Xuejun; Yang, Shaomin; Wang, Dijin</p> <p>2017-02-01</p> <p>In this study, we present a new synthesis of GPS <span class="hlt">velocities</span> for tectonic deformation within the Tibetan Plateau and its surrounding areas, a combined data set of ˜1854 GPS-derived horizontal <span class="hlt">velocity</span> vectors. Assuming that crustal deformation is localized along major faults, a block modelling approach is employed to interpret the GPS <span class="hlt">velocity</span> <span class="hlt">field</span>. We construct a 30-element block model to describe present-day deformation in western China, with half of them located within the Tibetan Plateau, and the remainder located in its surrounding areas. We model the GPS <span class="hlt">velocities</span> simultaneously for the effects of block rotations and elastic strain induced by the bounding faults. Our model yields a good fit to the GPS data with a mean residual of 1.08 mm a-1 compared to the mean uncertainty of 1.36 mm a-1 for each <span class="hlt">velocity</span> component, indicating a good agreement between the predicted and <span class="hlt">observed</span> <span class="hlt">velocities</span>. The major strike-slip faults such as the Altyn Tagh, Xianshuihe, Kunlun and Haiyuan faults have relatively uniform slip rates in a range of 5-12 mm a-1 along most of their segments, and the estimated fault slip rates agree well with previous geologic and geodetic results. Blocks having significant residuals are located at the southern and southeastern Tibetan Plateau, suggesting complex tectonic settings and further refinement of accurate definition of block geometry in these regions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.S41A4439C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.S41A4439C"><span>Shear <span class="hlt">velocity</span> of the Rotokawa geothermal <span class="hlt">field</span> using ambient noise</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Civilini, F.; Savage, M. K.; Townend, J.</p> <p>2014-12-01</p> <p>Ambient noise correlation is an increasingly popular seismological technique that uses the ambient seismic noise recorded at two stations to construct an empirical Green's function. Applications of this technique include determining shear <span class="hlt">velocity</span> structure and attenuation. An advantage of ambient noise is that it does not rely on external sources of seismic energy such as local or teleseismic earthquakes. This method has been used in the geothermal industry to determine the depths at which magmatic processes occur, to distinguish between production and non-production areas, and to <span class="hlt">observe</span> seismic <span class="hlt">velocity</span> perturbations associated with fluid extraction. We will present a <span class="hlt">velocity</span> model for the Rotokawa geothermal <span class="hlt">field</span> near Taupo, New Zealand, produced from ambient noise cross correlations. Production at Rotokawa is based on the "Rotokawa A" combined cycle power station established in 1997 and the "Nga Awa Purua" triple flash power plant established in 2010. Rotokawa Joint Venture, a partnership between Mighty River Power and Tauhara North No. 2 Trust currently operates 174 MW of generation at Rotokawa. An array of short period seismometers was installed in 2008 and occupies an area of roughly 5 square kilometers around the site. Although both cultural and natural noise sources are recorded at the stations, the instrument separation distance provides a unique challenge for analyzing cross correlations produced by both signal types. The inter-station spacing is on the order of a few kilometers, so waves from cultural sources generally are not coherent from one station to the other, while the wavelength produced by natural noise is greater than the station separation. <span class="hlt">Velocity</span> models produced from these two source types will be compared to known geological models of the site. Depending on the amount of data needed to adequately construct cross-correlations, a time-dependent model of <span class="hlt">velocity</span> will be established and compared with geothermal production processes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMEP43A0953M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMEP43A0953M"><span>Surprises from the <span class="hlt">field</span>: Novel aspects of aeolian saltation <span class="hlt">observed</span> under natural turbulence</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Martin, R. L.; Kok, J. F.; Chamecki, M.</p> <p>2015-12-01</p> <p>The mass flux of aeolian (wind-blown) sediment transport - critical for understanding earth and planetary geomorphology, dust generation, and soil stability - is difficult to predict. Recent work suggests that competing models for saltation (the characteristic hopping of aeolian sediment) fail because they do not adequately account for wind turbulence. To address this issue, we performed <span class="hlt">field</span> deployments measuring high-frequency co-variations of aeolian saltation and near-surface winds at multiple sites under a range of conditions. Our <span class="hlt">observations</span> yield several novel findings not currently captured by saltation models: (1) Saltation flux displays no significant lag relative to horizontal wind <span class="hlt">velocity</span>; (2) Characteristic height of the saltation layer remains constant with changes in shear <span class="hlt">velocity</span>; and (3) During saltation, the vertical profile of mean horizontal wind <span class="hlt">velocity</span> is steeper than expected from the Reynolds stress. We examine how the interactions between saltation and turbulence in <span class="hlt">field</span> settings could explain some of these surprising <span class="hlt">observations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014BTSNU..51...53K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014BTSNU..51...53K"><span>Monte-Carlo Method Application for Precising Meteor <span class="hlt">Velocity</span> from TV <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>Kozak, P.</p> <p>2014-12-01</p> <p>Monte-Carlo method (method of statistical trials) as an application for meteor <span class="hlt">observations</span> 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 <span class="hlt">velocity</span> of a meteor - which has the highest influence onto precision of meteor heliocentric orbit elements calculation. In classical approach the <span class="hlt">velocity</span> 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 <span class="hlt">observational</span> points. Then we calculated the absolute value of <span class="hlt">velocity</span> independently from each <span class="hlt">observational</span> point selecting any of them from some reasons as a final parameter. In the given method we propose to obtain a statistical distribution of <span class="hlt">velocity</span> absolute value as an intersection of two distributions corresponding to <span class="hlt">velocity</span> values obtained from different points. We suppose that such an approach has to substantially increase the precision of meteor <span class="hlt">velocity</span> calculation and remove any subjective inaccuracies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998A%26A...338.1041L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998A%26A...338.1041L"><span>Detection of atmospheric <span class="hlt">velocity</span> <span class="hlt">fields</span> in A-type stars</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Landstreet, J. D.</p> <p>1998-10-01</p> <p>High signal-to-noise spectra with spectral resolution of more than 10(5) have been obtained of one normal B9.5V, one normal A1V, two Am stars, and two HgMn B stars having v sin i less than 6 km s(-1) . These spectra are modeled with LTE line profile synthesis to test the extent to which the spectrum of each star can be modeled correctly with a single set of parameters T_e, log g, chemical abundances, v sin i, and (depth-independent) microturbulent <span class="hlt">velocity</span> xi . The answer to this question is important for abundance analysis of A and B stars; if conventional line synthesis does not reproduce the line profiles <span class="hlt">observed</span> in stars of small v sin i, results obtained from such analysis are not likely to be very precise. The comparison of models with <span class="hlt">observations</span> is then used to search for direct evidence of atmospheric motions, including line-strength dependent broadening, line core shape, and line asymmetries, in order to study how the microturbulence derived from abundance analysis is related to more direct evidence of atmospheric <span class="hlt">velocity</span> <span class="hlt">fields</span>. It is found for the three stars with 12,000 >= T_e >= 10,200 K (the normal star 21 Peg and the two HgMn stars 53 Tau and HD 193452) that xi is less than 1 km s(-1) , and line profiles are reproduced accurately by the synthesis with a single set of parameters. The slightly cooler (T_e ~ 9800 K) star HD 72660 has only a slightly stronger surface convective layer than the hotter stars, but for this star xi ~ 2.2 km s(-1) . Strong spectral lines all show significant asymmetry, with the blue line wing deeper than the red wing, and have line bisectors which have curvature towards the blue with a span of about 0.5 to 1.0 km s(-1) . A single model fits all lines satisfactorily. The two Am stars (HD 108642 and 32 Aqr), with T_e ~ 8000 K, are found to have much larger values of xi (4 to 5 km s(-1) ). The strong spectral lines of these two stars are extremely asymmetric, with depressed blue wings, and the bisectors have spans of order 3</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017OptEn..56e4102C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017OptEn..56e4102C"><span><span class="hlt">Field</span>-programmable gate array-controlled sweep <span class="hlt">velocity</span>-locked laser pulse generator</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Zhen; Hefferman, Gerald; Wei, Tao</p> <p>2017-05-01</p> <p>A <span class="hlt">field</span>-programmable gate array (FPGA)-controlled sweep <span class="hlt">velocity</span>-locked laser pulse generator (SV-LLPG) design based on an all-digital phase-locked loop (ADPLL) is proposed. A distributed feedback laser with modulated injection current was used as a swept-frequency laser source. An open-loop predistortion modulation waveform was calibrated using a feedback iteration method to initially improve frequency sweep linearity. An ADPLL control system was then implemented using an FPGA to lock the output of a Mach-Zehnder interferometer that was directly proportional to laser sweep <span class="hlt">velocity</span> to an on-board system clock. Using this system, linearly chirped laser pulses with a sweep bandwidth of 111.16 GHz were demonstrated. Further testing evaluating the sensing utility of the system was conducted. In this test, the SV-LLPG served as the swept laser source of an optical frequency-domain reflectometry system used to interrogate a subterahertz range fiber structure (sub-THz-FS) array. A static strain test was then conducted and linear sensor results were <span class="hlt">observed</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.S41B2753T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.S41B2753T"><span>Temporal Variability in Seismic <span class="hlt">Velocity</span> at the Salton Sea Geothermal <span class="hlt">Field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Taira, T.; Nayak, A.; Brenguier, F.</p> <p>2015-12-01</p> <p>We characterize the temporal variability of ambient noise wavefield and search for <span class="hlt">velocity</span> changes associated with activities of the geothermal energy development at the Salton Sea Geothermal <span class="hlt">Field</span>. The noise cross-correlations (NCFs) are computed for ~6 years of continuous three-component seismic data (December 2007 through January 2014) collected at 8 sites from the CalEnergy Subnetwork (EN network) with MSNoise software (Lecocq et al., 2014, SRL). All seismic data are downloaded from the Southern California Earthquake Data Center. <span class="hlt">Velocity</span> changes (dv/v) are obtained by measuring time delay between 5-day stacks of NCFs and the reference NCF (average over the entire 6 year period). The time history of dv/v is determined by averaging dv/v measurements over all station/channel pairs (252 combinations). Our preliminary dv/v measurement suggests a gradual increase in dv/v over the 6-year period in a frequency range of 0.5-8.0 Hz. The resultant increase rate of <span class="hlt">velocity</span> is about 0.01%/year. We also explore the frequency-dependent <span class="hlt">velocity</span> change at the 5 different frequency bands (0.5-2.0 Hz, 0.75-3.0 Hz, 1.0-4.0 Hz, 1.5-6.0 Hz, and 2.0-8.0 Hz) and find that the level of this long-term dv/v variability is increased with increase of frequency (i.e., the highest increase rate of ~0.15%/year at the 0.5-2.0 Hz band). This result suggests that the <span class="hlt">velocity</span> changes were mostly occurred in a depth of ~500 m assuming that the coda parts of NCFs (~10-40 s depending on station distances) are predominantly composed of scattered surface waves, with the SoCal <span class="hlt">velocity</span> model (Dreger and Helmberger, 1993, JGR). No clear seasonal variation of dv/v is <span class="hlt">observed</span> in the frequency band of 0.5-8.0 Hz.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140000913','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140000913"><span>Large-scale <span class="hlt">Observations</span> of a Subauroral Polarization Stream by Midlatitude SuperDARN Radars: Instantaneous Longitudinal <span class="hlt">Velocity</span> Variations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>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.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20140000913'); toggleEditAbsImage('author_20140000913_show'); toggleEditAbsImage('author_20140000913_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20140000913_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20140000913_hide"></p> <p>2012-01-01</p> <p>We present simultaneous measurements of flow <span class="hlt">velocities</span> inside a subauroral polarization stream (SAPS) made by six midlatitude high-frequency SuperDARN radars. The instantaneous <span class="hlt">observations</span> cover three hours of universal time and six hours of magnetic local time (MLT). From <span class="hlt">velocity</span> variations across the <span class="hlt">field</span>-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 <span class="hlt">velocity</span> on magnetic local time. We find that the SAPS <span class="hlt">velocity</span> variation with magnetic local time is best described by an exponential function. The average <span class="hlt">velocity</span> 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 <span class="hlt">observations</span> in good agreement with results from ring current models. Using TEC measurements we find that the high westward <span class="hlt">velocities</span> of the SAPS are - as expected - located in a region of low TEC values, indicating low ionospheric conductivities.</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_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li class="active"><span>5</span></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_5 --> <div id="page_6" 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_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</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="101"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFDA17004D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFDA17004D"><span>A potential method for lift evaluation from <span class="hlt">velocity</span> <span class="hlt">field</span> data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>de Guyon-Crozier, Guillaume; Mulleners, Karen</p> <p>2017-11-01</p> <p>Computing forces from <span class="hlt">velocity</span> <span class="hlt">field</span> measurements is one of the challenges in experimental aerodynamics. This work focuses on low Reynolds flows, where the dynamics of the leading and trailing edge vortices play a major role in lift production. Recent developments in 2D potential flow theory, using discrete vortex models, have shown good results for unsteady wing motions. A method is presented to calculate lift from experimental <span class="hlt">velocity</span> <span class="hlt">field</span> data using a discrete vortex potential flow model. The model continuously adds new point vortices at leading and trailing edges whose circulations are set directly from vorticity measurements. Forces are computed using the unsteady Blasius equation and compared with measured loads.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19920072497&hterms=Electric+current&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DElectric%2Bcurrent','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19920072497&hterms=Electric+current&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DElectric%2Bcurrent"><span>Correlation between magnetic and electric <span class="hlt">field</span> perturbations in the <span class="hlt">field</span>-aligned current regions deduced from DE 2 <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>Ishii, M.; Sugiura, M.; Iyemori, T.; Slavin, J. A.</p> <p>1992-01-01</p> <p>The satellite-<span class="hlt">observed</span> high correlations between magnetic and electric <span class="hlt">field</span> perturbations in the high-latitude <span class="hlt">field</span>-aligned current regions are investigated by examining the dependence of the relationship between Delta-B and E on spatial scale, using the electric and magnetic <span class="hlt">field</span> data obtained by DE 2 in the polar regions. The results are compared with the Pedersen conductivity inferred from the international reference ionosphere model and the Alfven wave <span class="hlt">velocity</span> calculated from the in situ ion density and magnetic <span class="hlt">field</span> measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFD.F4009V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFD.F4009V"><span>On magnetic <span class="hlt">field</span> strength effect on <span class="hlt">velocity</span> and turbulence characterization using Phase-Contrast Magnetic Resonance Imaging (PC-MRI)</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 Moortele, Pierre-Francois; Amili, Omid; Coletti, Filippo; Toloui, Mostafa</p> <p>2017-11-01</p> <p>Cardiovascular flows are predominantly laminar. Nevertheless, transient and even turbulent flows have been <span class="hlt">observed</span> in the vicinity of the heart (e.g. valves, ascending aorta, valvular/vascular stenosis). Effective in-vivo hemodynamic-based diagnostics in these sites require both high-resolution <span class="hlt">velocity</span> measurements (especially in the near-vessel wall regions) and accurate evaluation of blood flow turbulence level (e.g. in terms of TKE). In addition to phase contrast (PC), appropriately designed PC-MRI sequences provide intravoxel incoherent motion encoding, a unique tool for simultaneous, non-invasive evaluation of <span class="hlt">velocity</span> 3D vector <span class="hlt">fields</span> and Reynolds stresses in cardiovascular flows in vivo. However, limited spatial and temporal resolution of PC-MRI result in inaccuracies in the estimation of hemodynamics (e.g. WSS) and of flow turbulence characteristics. This study aims to assess whether SNR gains at higher magnetic <span class="hlt">field</span> could overcome these limits, providing more accurate <span class="hlt">velocity</span> and turbulence characterization at higher spatial resolution. Experiments are conducted on MR Scanners at 3 and 7 Tesla with a U-bent pipe flow shaped phantom. 3D <span class="hlt">velocity</span> <span class="hlt">fields</span>, Reynolds stresses and TKE are analyzed and compared to a reference PIV experiments.</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 <span class="hlt">FIELDS</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>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 <span class="hlt">fields</span> 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 gradient, the median [Fe/H] also declines with increasing Galactic latitude in off-axis <span class="hlt">fields</span>. The off-axis vertical [Fe/H] gradient 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 gradient does not exist. The (+8.5, +9) <span class="hlt">field</span> 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) <span class="hlt">field</span>. However, the overall radial <span class="hlt">velocity</span> and dispersion for each <span class="hlt">field</span> are in good agreement with recent surveys and bulge models. All <span class="hlt">fields</span> 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 <span class="hlt">fields</span>. Additionally, the [O/Fe], [Si</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21460080-velocity-field-compressible-magnetohydrodynamic-turbulence-wavelet-decomposition-mode-scalings','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21460080-velocity-field-compressible-magnetohydrodynamic-turbulence-wavelet-decomposition-mode-scalings"><span><span class="hlt">VELOCITY</span> <span class="hlt">FIELD</span> OF COMPRESSIBLE MAGNETOHYDRODYNAMIC TURBULENCE: WAVELET DECOMPOSITION AND MODE SCALINGS</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>Kowal, Grzegorz; Lazarian, A., E-mail: kowal@astro.wisc.ed, E-mail: lazarian@astro.wisc.ed</p> <p></p> <p>We study compressible magnetohydrodynamic turbulence, which holds the key to many astrophysical processes, including star formation and cosmic-ray propagation. To account for the variations of the magnetic <span class="hlt">field</span> in the strongly turbulent fluid, we use wavelet decomposition of the turbulent <span class="hlt">velocity</span> <span class="hlt">field</span> into Alfven, slow, and fast modes, which presents an extension of the Cho and Lazarian decomposition approach based on Fourier transforms. The wavelets allow us to follow the variations of the local direction of the magnetic <span class="hlt">field</span> and therefore improve the quality of the decomposition compared to the Fourier transforms, which are done in the mean <span class="hlt">field</span> referencemore » frame. For each resulting component, we calculate the spectra and two-point statistics such as longitudinal and transverse structure functions as well as higher order intermittency statistics. In addition, we perform a Helmholtz- Hodge decomposition of the <span class="hlt">velocity</span> <span class="hlt">field</span> into incompressible and compressible parts and analyze these components. We find that the turbulence intermittency is different for different components, and we show that the intermittency statistics depend on whether the phenomenon was studied in the global reference frame related to the mean magnetic <span class="hlt">field</span> or in the frame defined by the local magnetic <span class="hlt">field</span>. The dependencies of the measures we obtained are different for different components of the <span class="hlt">velocity</span>; for instance, we show that while the Alfven mode intermittency changes marginally with the Mach number, the intermittency of the fast mode is substantially affected by the change.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018EP%26S...70....1K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018EP%26S...70....1K"><span>Difference of horizontal-to-vertical spectral ratios of <span class="hlt">observed</span> earthquakes and microtremors and its application to S-wave <span class="hlt">velocity</span> inversion based on the diffuse <span class="hlt">field</span> concept</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kawase, Hiroshi; Mori, Yuta; Nagashima, Fumiaki</p> <p>2018-01-01</p> <p>We have been discussing the validity of using the horizontal-to-vertical spectral ratios (HVRs) as a substitute for S-wave amplifications after Nakamura first proposed the idea in 1989. So far a formula for HVRs had not been derived that fully utilized their physical characteristics until a recent proposal based on the diffuse <span class="hlt">field</span> concept. There is another source of confusion that comes from the mixed use of HVRs from earthquake and microtremors, although their wave <span class="hlt">fields</span> are hardly the same. In this study, we compared HVRs from <span class="hlt">observed</span> microtremors (MHVR) and those from <span class="hlt">observed</span> earthquake motions (EHVR) at one hundred K-NET and KiK-net stations. We found that MHVR and EHVR share similarities, especially until their first peak frequency, but have significant differences in the higher frequency range. This is because microtremors mainly consist of surface waves so that peaks associated with higher modes would not be prominent, while seismic motions mainly consist of upwardly propagating plain body waves so that higher mode resonances can be seen in high frequency. We defined here the spectral amplitude ratio between them as EMR and calculated their average. We categorize all the sites into five bins by their fundamental peak frequencies in MHVR. Once we obtained EMRs for five categories, we back-calculated EHVRs from MHVRs, which we call pseudo-EHVRs (pEHVR). We found that pEHVR is much closer to EHVR than MHVR. Then we use our inversion code to invert the one-dimensional S-wave <span class="hlt">velocity</span> structures from EHVRs based on the diffuse <span class="hlt">field</span> concept. We also applied the same code to pEHVRs and MHVRs for comparison. We found that pEHVRs yield <span class="hlt">velocity</span> structures much closer to those by EHVRs than those by MHVRs. This is natural since what we have done up to here is circular except for the average operation in EMRs. Finally, we showed independent examples of data not used in the EMR calculation, where better ground structures were successfully identified from p</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JBO....23b6008W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JBO....23b6008W"><span>Wide-<span class="hlt">field</span> absolute transverse blood flow <span class="hlt">velocity</span> mapping in vessel centerline</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Nanshou; Wang, Lei; Zhu, Bifeng; Guan, Caizhong; Wang, Mingyi; Han, Dingan; Tan, Haishu; Zeng, Yaguang</p> <p>2018-02-01</p> <p>We propose a wide-<span class="hlt">field</span> absolute transverse blood flow <span class="hlt">velocity</span> measurement method in vessel centerline based on absorption intensity fluctuation modulation effect. The difference between the light absorption capacities of red blood cells and background tissue under low-coherence illumination is utilized to realize the instantaneous and average wide-<span class="hlt">field</span> optical angiography images. The absolute fuzzy connection algorithm is used for vessel centerline extraction from the average wide-<span class="hlt">field</span> optical angiography. The absolute transverse <span class="hlt">velocity</span> in the vessel centerline is then measured by a cross-correlation analysis according to instantaneous modulation depth signal. The proposed method promises to contribute to the treatment of diseases, such as those related to anemia or thrombosis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008JPhCS.135a2020B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008JPhCS.135a2020B"><span>Methodology to estimate the relative pressure <span class="hlt">field</span> from noisy experimental <span class="hlt">velocity</span> data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bolin, C. D.; Raguin, L. G.</p> <p>2008-11-01</p> <p>The determination of intravascular pressure <span class="hlt">fields</span> is important to the characterization of cardiovascular pathology. We present a two-stage method that solves the inverse problem of estimating the relative pressure <span class="hlt">field</span> from noisy <span class="hlt">velocity</span> <span class="hlt">fields</span> measured by phase contrast magnetic resonance imaging (PC-MRI) on an irregular domain with limited spatial resolution, and includes a filter for the experimental noise. For the pressure calculation, the Poisson pressure equation is solved by embedding the irregular flow domain into a regular domain. To lessen the propagation of the noise inherent to the <span class="hlt">velocity</span> measurements, three filters - a median filter and two physics-based filters - are evaluated using a 2-D Couette flow. The two physics-based filters outperform the median filter for the estimation of the relative pressure <span class="hlt">field</span> for realistic signal-to-noise ratios (SNR = 5 to 30). The most accurate pressure <span class="hlt">field</span> results from a filter that applies in a least-squares sense three constraints simultaneously: consistency between measured and filtered <span class="hlt">velocity</span> <span class="hlt">fields</span>, divergence-free and additional smoothness conditions. This filter leads to a 5-fold gain in accuracy for the estimated relative pressure <span class="hlt">field</span> compared to without noise filtering, in conditions consistent with PC-MRI of the carotid artery: SNR = 5, 20 x 20 discretized flow domain (25 X 25 computational domain).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930039175&hterms=angular+velocity&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dangular%2Bvelocity','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930039175&hterms=angular+velocity&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dangular%2Bvelocity"><span>Inference of the angular <span class="hlt">velocity</span> of plasma in the Jovian magnetosphere from the sweepback of magnetic <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Khurana, Krishan K.; Kivelson, Margaret G.</p> <p>1993-01-01</p> <p>The averaged angular <span class="hlt">velocity</span> of plasma from magnetic <span class="hlt">observations</span> is evaluated using plasma outflow rate as a parameter. New techniques are developed to calculate the normal and azimuthal components of the magnetic <span class="hlt">field</span> in and near to the plasma sheet in a plasma sheet coordinate system. The revised <span class="hlt">field</span> components differ substantially from the quantities used in previous analyses. With the revised <span class="hlt">field</span> values, it appears that during the Voyager 2 flyby for an outflow rate of 2.5 x 10 exp 29 amu/s, the <span class="hlt">observed</span> magnetic torque may be sufficient to keep the plasma in corotation to radial distances of 50 Rj in the postmidnight quadrant.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26628182','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26628182"><span>Note: Device for obtaining volumetric, three-component <span class="hlt">velocity</span> <span class="hlt">fields</span> inside cylindrical cavities.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ramírez, G; Núñez, J; Hernández, G N; Hernández-Cruz, G; Ramos, E</p> <p>2015-11-01</p> <p>We describe a device designed and built to obtain the three-component, steady state <span class="hlt">velocity</span> <span class="hlt">field</span> in the whole volume occupied by a fluid in motion contained in a cavity with cylindrical walls. The prototype comprises a two-camera stereoscopic particle image velocimetry system mounted on a platform that rotates around the volume under analysis and a slip ring arrangement that transmits data from the rotating sensors to the data storage elements. Sample <span class="hlt">observations</span> are presented for natural convection in a cylindrical container but other flows can be analyzed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19960008425','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19960008425"><span>Full <span class="hlt">field</span> gas phase <span class="hlt">velocity</span> measurements in microgravity</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Griffin, Devon W.; Yanis, William</p> <p>1995-01-01</p> <p>Measurement of full-<span class="hlt">field</span> <span class="hlt">velocities</span> via Particle Imaging Velocimetry (PIV) is common in research efforts involving fluid motion. While such measurements have been successfully performed in the liquid phase in a microgravity environment, gas-phase measurements have been beset by difficulties with seeding and laser strength. A synthesis of techniques developed at NASA LeRC exhibits promise in overcoming these difficulties. Typical implementation of PIV involves forming the light from a pulsed laser into a sheet that is some fraction of a millimeter thick and 50 or more millimeters wide. When a particle enters this sheet during a pulse, light scattered from the particle is recorded by a detector, which may be a film plane or a CCD array. Assuming that the particle remains within the boundaries of the sheet for the second pulse and can be distinguished from neighboring particles, comparison of the two images produces an average <span class="hlt">velocity</span> vector for the time between the pulses. If the concentration of particles in the sampling volume is sufficiently large but the particles remain discrete, a full <span class="hlt">field</span> map may be generated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110023043','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110023043"><span>Sodium Atoms in the Lunar Exotail: <span class="hlt">Observed</span> <span class="hlt">Velocity</span> and Spatial Distributions</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Line, Michael R.; Mierkiewicz, E. J.; Oliversen, R. J.; Wilson, J. K.; Haffner, L. M.; Roesler, F. L.</p> <p>2011-01-01</p> <p>The lunar sodium tail extends long distances due to radiation pressure on sodium atoms in the lunar exosphere. Our earlier <span class="hlt">observations</span> determined the average radial <span class="hlt">velocity</span> 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 <span class="hlt">velocity</span> 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 <span class="hlt">observations</span> 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 <span class="hlt">velocity</span> distributions in the sodium exotail are sensitive to the near surface lunar sodium <span class="hlt">velocity</span> distribution and that <span class="hlt">observations</span> of this sort along with detailed modeling offer new opportunities to describe the time history of lunar surface sputtering over several days.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4783027','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4783027"><span>Global Neuromagnetic Cortical <span class="hlt">Fields</span> Have Non-Zero <span class="hlt">Velocity</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>Alexander, David M.; Nikolaev, Andrey R.; Jurica, Peter; Zvyagintsev, Mikhail; Mathiak, Klaus; van Leeuwen, Cees</p> <p>2016-01-01</p> <p>Globally coherent patterns of phase can be obscured by analysis techniques that aggregate brain activity measures across-trials, whether prior to source localization or for estimating inter-areal coherence. We analyzed, at single-trial level, whole head MEG recorded during an <span class="hlt">observer</span>-triggered apparent motion task. Episodes of globally coherent activity occurred in the delta, theta, alpha and beta bands of the signal in the form of large-scale waves, which propagated with a variety of <span class="hlt">velocities</span>. Their mean speed at each frequency band was proportional to temporal frequency, giving a range of 0.06 to 4.0 m/s, from delta to beta. The wave peaks moved over the entire measurement array, during both ongoing activity and task-relevant intervals; direction of motion was more predictable during the latter. A large proportion of the cortical signal, measurable at the scalp, exists as large-scale coherent motion. We argue that the distribution of <span class="hlt">observable</span> phase <span class="hlt">velocities</span> in MEG is dominated by spatial filtering considerations in combination with group <span class="hlt">velocity</span> of cortical activity. Traveling waves may index processes involved in global coordination of cortical activity. PMID:26953886</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> <span class="hlt">field</span> 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> <span class="hlt">field</span> 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> gradient 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 <span class="hlt">field</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19730014251','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19730014251"><span><span class="hlt">Observations</span> of disk-shaped bodies in free flight at terminal <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>Vorreiter, J. W.; Tate, D. L.</p> <p>1973-01-01</p> <p>Ten disk-shaped models of a proposed nuclear heat source module were released from an aircraft and <span class="hlt">observed</span> by radar. The initial launch attitude, spin rate, and mass of the models were varied. Significant differences were <span class="hlt">observed</span> in the mode of flight and terminal <span class="hlt">velocity</span> 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 <span class="hlt">velocity</span> of the models was found to be well correlated as a function of mass per unit frontal area. The demonstrated terminal <span class="hlt">velocity</span> of the modular disk heat source, about 27 m/sec for this specific design, is only 33% of that of existing heat source designs.</p> </li> <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 <span class="hlt">field</span>. 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/2014AGUFM.G31A0388F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.G31A0388F"><span>Uncertainty of InSAR <span class="hlt">velocity</span> <span class="hlt">fields</span> for measuring long-wavelength displacement</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fattahi, H.; Amelung, F.</p> <p>2014-12-01</p> <p>Long-wavelength artifacts in InSAR data are the main limitation to measure long-wavelength displacement; they are traditionally attributed mainly to the inaccuracy of the satellite orbits (orbital errors). However, most satellites are precisely tracked resulting in uncertainties of orbits of 2-10 cm. Orbits of these satellites are thus precise enough to obtain precise <span class="hlt">velocity</span> <span class="hlt">fields</span> with uncertainties better than 1 mm/yr/100 km for older satellites (e.g. Envisat) and better than 0.2 mm/yr/100 km for modern satellites (e.g. TerraSAR-X and Sentinel-1) [Fattahi & Amelung, 2014]. Such accurate <span class="hlt">velocity</span> <span class="hlt">fields</span> are achievable if long-wavelength artifacts from sources other than orbital errors are identified and corrected for. We present a modified Small Baseline approach to measure long-wavelength deformation and evaluate the uncertainty of these measurements. We use a redundant network of interferograms for detection and correction of unwrapping errors to ensure the unbiased estimation of phase history. We distinguish between different sources of long-wavelength artifacts and correct those introduced by atmospheric delay, topographic residuals, timing errors, processing approximations and hardware issues. We evaluate the uncertainty of the <span class="hlt">velocity</span> <span class="hlt">fields</span> using a covariance matrix with the contributions from orbital errors and residual atmospheric delay. For contributions from the orbital errors we consider the standard deviation of <span class="hlt">velocity</span> gradients in range and azimuth directions as a function of orbital uncertainty. For contributions from the residual atmospheric delay we use several approaches including the structure functions of InSAR time-series epochs, the predicted delay from numerical weather models and estimated wet delay from optical imagery. We validate this InSAR approach for measuring long-wavelength deformation by comparing InSAR <span class="hlt">velocity</span> <span class="hlt">fields</span> over ~500 km long swath across the southern San Andreas fault system with independent GPS <span class="hlt">velocities</span> and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H34A..04O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H34A..04O"><span><span class="hlt">Field</span> Testing of an In-well Point <span class="hlt">Velocity</span> Probe for the Rapid Characterization of 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>Osorno, T.; Devlin, J. F.</p> <p>2017-12-01</p> <p>Reliable estimates of groundwater <span class="hlt">velocity</span> is essential in order to best implement in-situ monitoring and remediation technologies. The In-well Point <span class="hlt">Velocity</span> Probe (IWPVP) is an inexpensive, reusable tool developed for rapid measurement of groundwater <span class="hlt">velocity</span> at the centimeter-scale in monitoring wells. IWPVP measurements of groundwater speed are based on a small-scale tracer test conducted as ambient groundwater passes through the well screen and the body of the probe. Horizontal flow direction can be determined from the difference in tracer mass passing detectors placed in four funnel-and-channel pathways through the probe, arranged in a cross pattern. The design viability of the IWPVP was confirmed using a two-dimensional numerical model in Comsol Multiphysics, followed by a series of laboratory tank experiments in which IWPVP measurements were calibrated to quantify seepage <span class="hlt">velocities</span> in both fine and medium sand. Lab results showed that the IWPVP was capable of measuring the seepage <span class="hlt">velocity</span> in less than 20 minutes per test, when the seepage <span class="hlt">velocity</span> was in the range of 0.5 to 4.0 m/d. Further, the IWPVP estimated the groundwater speed with a precision of ± 7%, and an accuracy of ± 14%, on average. The horizontal flow direction was determined with an accuracy of ± 15°, on average. Recently, a pilot <span class="hlt">field</span> test of the IWPVP was conducted in the Borden aquifer, C.F.B. Borden, Ontario, Canada. A total of approximately 44 IWPVP tests were conducted within two 2-inch groundwater monitoring wells comprising a 5 ft. section of #8 commercial well screen. Again, all tests were completed in under 20 minutes. The <span class="hlt">velocities</span> estimated from IWPVP data were compared to 21 Point <span class="hlt">Velocity</span> Probe (PVP) tests, as well as Darcy-based estimates of groundwater <span class="hlt">velocity</span>. Preliminary data analysis shows strong agreement between the IWPVP and PVP estimates of groundwater <span class="hlt">velocity</span>. Further, both the IWPVP and PVP estimates of groundwater <span class="hlt">velocity</span> appear to be reasonable when</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017GeoRL..4411942Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GeoRL..4411942Z"><span>Propagation of the Semidiurnal Internal Tide: Phase <span class="hlt">Velocity</span> Versus Group <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>Zhao, Zhongxiang</p> <p>2017-12-01</p> <p>The superposition of two waves of slightly different wavelengths has long been used to illustrate the distinction between phase <span class="hlt">velocity</span> and group <span class="hlt">velocity</span>. 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 <span class="hlt">velocity</span>. Long-range propagation of M2 and S2 internal tides in the North Pacific is <span class="hlt">observed</span> 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 <span class="hlt">velocity</span>, the S2 phase <span class="hlt">velocity</span>, and the group <span class="hlt">velocity</span> are determined from phase gradients of the corresponding satellite <span class="hlt">observed</span> internal tide <span class="hlt">fields</span>. The <span class="hlt">observed</span> phase and group <span class="hlt">velocities</span> agree with theoretical values estimated using the World Ocean Atlas 2013 annual-mean ocean stratification.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008JLTP..151.1113S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008JLTP..151.1113S"><span>Sound <span class="hlt">Velocity</span> Measurements in the Low and the High <span class="hlt">Field</span> Phases of the Nuclear-Ordered bcc Solid 3He in Magnetic <span class="hlt">Fields</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sasaki, Satoshi; Nakayama, Atsuyoshi; Sasaki, Yutaka; Mizusaki, Takao</p> <p>2008-06-01</p> <p>We have measured the temperature and magnetic-<span class="hlt">field</span> dependences of the sound <span class="hlt">velocity</span> for one longitudinal and two transverse waves in the low <span class="hlt">field</span> phase (LFP) and the high <span class="hlt">field</span> phase (HFP) of nuclear spin ordered bcc solid 3He crystals with a single magnetic domain along the melting curve. From sound <span class="hlt">velocity</span> measurements for various crystal orientations as a function of the sound propagation direction, we determined the elastic stiffness constants, c ij ( T, B). In the LFP with tetragonal symmetry for the nuclear spin structure, we extracted six nuclear spin elastic stiffness constants Δ c {/ij ℓ }( T,0.06 T) from the temperature dependence of the sound <span class="hlt">velocity</span> at 0.06 T and Δ c {/ij ℓ }(0.5 mK, B) from the magnetic-<span class="hlt">field</span> dependence of sound <span class="hlt">velocity</span> at 0.5 mK. In the HFP with cubic symmetry for the nuclear spin structure, we extracted three Δ c {/ij h }( T,0.50 T) at 0.50 T and Δ c {/ij h }(0.5 mK, B) at 0.5 mK. At the first-order magnetic phase transition from the LFP to the HFP at the lower critical <span class="hlt">field</span> B c1, large jumps in sound <span class="hlt">velocities</span> were <span class="hlt">observed</span> for various crystal directions and we extracted three Δ c_{ij}^{total}|_{B_{c1}} . Using the thermodynamic relation between Δ c ij and the change in the internal energy for the exchange interaction in this system, Δ U ex( T, B), Δ c ij are related to the generalized second-order Grüneisen constants Γ{/ij X }≡ ∂ 2ln X/ ∂ ɛ i ∂ ɛ j as Δ c ij ( T, B)=Γ{/ij X }Δ U ex( T, B), where X represents some physical quantity which depends on the molar volume and ɛ j is the j-th component of a strain tensor. In the LFP, the Δ c {/ij ℓ }( T,0.06 T) were proportional to T 4, and Δ c {/ij ℓ }(0.5 mK, B) were proportional to B 2. We extracted Γ_{ij}^{s^{ell}} for the spin wave <span class="hlt">velocity</span> in the LFP, s ℓ , from Δ c {/ij ℓ }( T,0.06 T) and Γ^{1/χ^{ell}}_{ij} for the inverse susceptibility, 1/ χ ℓ from Δ c {/ij ℓ }(0.5 mK, B). In the HFP, Δ c {/ij h }( T,0.50 T) were proportional</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_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_6 --> <div id="page_7" 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_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> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="121"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/34854','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/34854"><span>Experimental investigation of the <span class="hlt">velocity</span> <span class="hlt">field</span> in buoyant diffusion flames using PIV and TPIV algorithm</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>L. Sun; X. Zhou; S.M. Mahalingam; D.R. Weise</p> <p>2005-01-01</p> <p>We investigated a simultaneous temporally and spatially resolved 2-D <span class="hlt">velocity</span> <span class="hlt">field</span> above a burning circular pan of alcohol using particle image velocimetry (PIV). The results obtained from PIV were used to assess a thermal particle image velocimetry (TPIV) algorithm previously developed to approximate the <span class="hlt">velocity</span> <span class="hlt">field</span> using the temperature <span class="hlt">field</span>, simultaneously...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSM43C..03K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSM43C..03K"><span>Superposed epoch analysis of vertical ion <span class="hlt">velocity</span>, electron temperature, <span class="hlt">field</span>-aligned current, and thermospheric wind in the dayside auroral region as <span class="hlt">observed</span> by DMSP and CHAMP</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kervalishvili, G.; Lühr, H.</p> <p>2016-12-01</p> <p>This study reports on the results obtained by a superposed epoch analysis (SEA) method applied to the electron temperature, vertical ion <span class="hlt">velocity</span>, <span class="hlt">field</span>-aligned current (FAC), and thermospheric zonal wind <span class="hlt">velocity</span> at high-latitudes in the Northern Hemisphere. The SEA study is performed in a magnetic latitude versus magnetic local time (MLat-MLT) frame. The obtained results are based on <span class="hlt">observations</span> collected during the years 2001-2005 by the CHAMP and DMSP (F13 and F15) satellites. The dependence on interplanetary magnetic <span class="hlt">field</span> (IMF) orientations is also investigated using data from the NASA/GSFC's OMNI database. Further, the obtained results are subdivided into three Lloyd seasons of 130 days each, which are defined as follows: local winter (1 January ± 65 days), combined equinoxes (1 April and 1 October ± 32days), and local summer (1 July ± 65 days). A period of 130 days is needed by the CHAMP satellite to pass through all local times. The time and location of the electron temperature peaks from CHAMP measurements near the cusp region are used as the reference parameter for the SEA method to investigate the relationship between the electron temperature and other ionospheric quantities. The SEA derived MLat profiles of the electron temperature show a seasonal dependence, increasing from winter to summer, as expected. But, the temperature rise (difference between the reference temperature peak and the background electron temperature) strongly decreases towards local summer. The SEA derived MLat profiles of the ion vertical <span class="hlt">velocity</span> at DMSP altitude show the same seasonal behaviour as the electron temperature rice. There exists a clear linear relation between these two variables with a quiet large correlation coefficient value, >0.9. The SEA derived MLat profiles of both, thermospheric zonal wind <span class="hlt">velocity</span> and FAC, show a clear IMF By orientation dependence for all local seasons. The zonal wind <span class="hlt">velocity</span> is prominently directed towards west in the MLat-MLT frame</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 <span class="hlt">field</span>-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 <span class="hlt">field</span>-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 <span class="hlt">field</span>-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 gradients 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/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://adsabs.harvard.edu/abs/2018PhFl...30c5107Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhFl...30c5107Y"><span>Effect of vortical structures on <span class="hlt">velocity</span> and turbulent <span class="hlt">fields</span> in the near region of an impinging turbulent jet</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yadav, Harekrishna; Agrawal, Amit</p> <p>2018-03-01</p> <p>This experimental study pertains to the formation of a secondary peak in heat transfer distribution for an axisymmetric turbulent impinging submerged jet. The analysis of instantaneous <span class="hlt">fields</span> is undertaken at various Reynolds numbers based upon the bulk <span class="hlt">velocity</span> and nozzle diameter (Re = 1300-10 000) and surface spacings (L/D = 0.25-6). Our analysis shows that flow separation and reattachment correspond to decrease/increase in local pressure and are caused by primary vortices; these are further linked to the location of maxima in streamwise and cross-stream <span class="hlt">velocities</span>. It is further <span class="hlt">observed</span> that the locations of maxima and minima in <span class="hlt">velocities</span> are linked to fluctuations in rms <span class="hlt">velocities</span> and thickening/thinning of the boundary layer. The vortices transported along the surface either coalesce among themselves or combine with other eddies to form a primary vortex. The primary vortex while getting convected downstream makes multiple interactions with the inner shear layer and causes waviness in instantaneous flow <span class="hlt">fields</span>. In their later stage, the primary vortex moves away from the wall and accelerates, while the flow decelerates in the inner shear layer. The accelerated fluid in the outer shear layer pulls the downstream fluid from the inner shear layer and leads to the formation of a secondary vortex. After a certain distance downstream, the secondary vortex rolling between the primary vortex and the wall eventually breaks down, while the flow reattaches to the wall. The behavior of time average and instantaneous <span class="hlt">velocity</span> <span class="hlt">fields</span> suggests that unsteadiness in the heat transfer is linked to the location of maximum streamwise <span class="hlt">velocity</span>, location of flow attachment, location of rms <span class="hlt">velocity</span>, and thickness of the boundary layer. The instantaneous <span class="hlt">velocity</span> <span class="hlt">fields</span> show that for a given surface spacing, the chances for the appearance of the secondary vortex reduce with an increase in Reynolds number because of the reduction in space available for the secondary vortex to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19740025365&hterms=channels+distribution&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dchannels%2Bdistribution','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19740025365&hterms=channels+distribution&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dchannels%2Bdistribution"><span>Channel flow analysis. [<span class="hlt">velocity</span> distribution throughout blade flow <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Katsanis, T.</p> <p>1973-01-01</p> <p>The design of a proper blade profile requires calculation of the blade row flow <span class="hlt">field</span> in order to determine the <span class="hlt">velocities</span> on the blade surfaces. An analysis theory is presented for several methods used for this calculation and associated computer programs that were developed are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.S43A2815C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.S43A2815C"><span>Temporal changes in shear <span class="hlt">velocity</span> from ambient noise at New Zealand geothermal <span class="hlt">fields</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Civilini, F.; Savage, M. K.; Townend, J.</p> <p>2016-12-01</p> <p>We use ambient noise to compare shear <span class="hlt">velocity</span> changes with geothermal production processes at the Ngatamariki and Rotokawa geothermal <span class="hlt">fields</span>, located in the central North Island of New Zealand. We calculate shear <span class="hlt">velocity</span> changes through an analysis of cross correlation functions of diffusive seismic wavefields between stations, which are proportional to Green's functions of the station path. Electricity production at Ngatamariki uses an 82 MW binary type power station manufactured by Ormat Technologies, which began operations in mid-2013 and is owned and operated by Mighty River Power. The "Nga Awa Purua" triple flash power plant at the Rotokawa geothermal <span class="hlt">field</span> was established in 2010 with parnership between Mighty River Power and Tauhara North No. 2 trust and currently operates 174 MW of generation. The seismometers of both networks, deployed primarily to <span class="hlt">observe</span> microseismicity within the <span class="hlt">field</span>, were installed prior to well stimulation and the start of production. Although cultural noise dominates the energy spectrum, a strong natural ambient noise signal can be detected when filtering below 1 Hz. Despite similar noise settings, the signal-to-noise ratio of cross correlation stacks at Rotokawa was more than two times greater than at Ngatamariki. We use stacks of cross correlations between stations prior to the onset of production as references, and compare them with cross correlations of moving stacks in time periods of well stimulation and the onset of electricity production.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19920045467&hterms=divided+attention&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Ddivided%2Battention','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19920045467&hterms=divided+attention&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Ddivided%2Battention"><span>Electric <span class="hlt">field</span> <span class="hlt">observations</span> of equatorial bubbles</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Aggson, T. L.; Maynard, N. C.; Hanson, W. B.; Saba, Jack L.</p> <p>1992-01-01</p> <p>Results from the double floating probe experiment performed on the San Marco D satellite are presented, with emphasis on the <span class="hlt">observation</span> of large incremental changes in the convective electric <span class="hlt">field</span> vector at the boundary of equatorial plasma bubbles. Attention is given to isolated bubble structures in the upper ionospheric F regions; these <span class="hlt">observed</span> bubble encounters are divided into two types - type I (live bubbles) and type II (dead bubbles). Type I bubbles show varying degrees of plasma depletion and large upward <span class="hlt">velocities</span> range up to 1000 km/s. The geometry of these bubbles is such that the spacecraft orbit may cut them where they are tilting either eastward or (more often) westward. Type II bubbles exhibit plasma density depletion but no appreciable upward convection. Both types of events are usually surrounded by a halo of plasma turbulence, which can extend considerably beyond the region of plasma depletion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19910002750','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19910002750"><span>Laser transit anemometer measurements of a JANNAF nozzle base <span class="hlt">velocity</span> flow <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hunter, William W., Jr.; Russ, C. E., Jr.; Clemmons, J. I., Jr.</p> <p>1990-01-01</p> <p><span class="hlt">Velocity</span> flow <span class="hlt">fields</span> of a nozzle jet exhausting into a supersonic flow were surveyed. The measurements were obtained with a laser transit anemometer (LTA) system in the time domain with a correlation instrument. The LTA data is transformed into the <span class="hlt">velocity</span> domain to remove the error that occurs when the data is analyzed in the time domain. The final data is shown in <span class="hlt">velocity</span> vector plots for positions upstream, downstream, and in the exhaust plane of the jet nozzle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4309966','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4309966"><span><span class="hlt">Observation</span> of Wave Packet Distortion during a Negative-Group-<span class="hlt">Velocity</span> Transmission</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ye, Dexin; Salamin, Yannick; Huangfu, Jiangtao; Qiao, Shan; Zheng, Guoan; Ran, Lixin</p> <p>2015-01-01</p> <p>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 <span class="hlt">velocity</span> 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 <span class="hlt">velocity</span>. 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-<span class="hlt">velocity</span> transmission of half-sine wave packets. We experimentally <span class="hlt">observe</span> the wave front and the distortion of modulated wave packets propagating with a negative group <span class="hlt">velocity</span> 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 <span class="hlt">observed</span>. This result agrees with Brillouin's assertion, i.e., the severe distortion of seemingly superluminal wave packets makes the definition of group <span class="hlt">velocity</span> physically meaningless in the anomalously dispersive region. PMID:25631746</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22522161-radial-velocity-variability-field-brown-dwarfs','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22522161-radial-velocity-variability-field-brown-dwarfs"><span>RADIAL <span class="hlt">VELOCITY</span> VARIABILITY OF <span class="hlt">FIELD</span> BROWN 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>Prato, L.; Mace, G. N.; Rice, E. L.</p> <p>2015-07-20</p> <p>We present paper six of the NIRSPEC Brown Dwarf Spectroscopic Survey, an analysis of multi-epoch, high-resolution (R ∼ 20,000) spectra of 25 <span class="hlt">field</span> dwarf systems (3 late-type M dwarfs, 16 L dwarfs, and 6 T dwarfs) taken with the NIRSPEC infrared spectrograph at the W. M. Keck Observatory. With a radial <span class="hlt">velocity</span> (RV) precision of ∼2 km s{sup −1}, we are sensitive to brown dwarf companions in orbits with periods of a few years or less given a mass ratio of 0.5 or greater. We do not detect any spectroscopic binary brown dwarfs in the sample. Given our target properties,more » and the frequency and cadence of <span class="hlt">observations</span>, we use a Monte Carlo simulation to determine the detection probability of our sample. Even with a null detection result, our 1σ upper limit for very low mass binary frequency is 18%. Our targets included seven known, wide brown dwarf binary systems. No significant RV variability was measured in our multi-epoch <span class="hlt">observations</span> of these systems, even for those pairs for which our data spanned a significant fraction of the orbital period. Specialized techniques are required to reach the high precisions sensitive to motion in orbits of very low-mass systems. For eight objects, including six T dwarfs, we present the first published high-resolution spectra, many with high signal to noise, that will provide valuable comparison data for models of brown dwarf atmospheres.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApJ...850...61Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApJ...850...61Y"><span>The Most Ancient Spiral Galaxy: A 2.6-Gyr-old Disk with a Tranquil <span class="hlt">Velocity</span> <span class="hlt">Field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yuan, Tiantian; Richard, Johan; Gupta, Anshu; Federrath, Christoph; Sharma, Soniya; Groves, Brent A.; Kewley, Lisa J.; Cen, Renyue; Birnboim, Yuval; Fisher, David B.</p> <p>2017-11-01</p> <p>We report an integral-<span class="hlt">field</span> spectroscopic (IFS) <span class="hlt">observation</span> of a gravitationally lensed spiral galaxy A1689B11 at redshift z = 2.54. It is the most ancient spiral galaxy discovered to date and the second kinematically confirmed spiral at z≳ 2. Thanks to gravitational lensing, this is also by far the deepest IFS <span class="hlt">observation</span> with the highest spatial resolution (˜400 pc) on a spiral galaxy at a cosmic time when the Hubble sequence is about to emerge. After correcting for a lensing magnification of 7.2 ± 0.8, this primitive spiral disk has an intrinsic star formation rate of 22 ± 2 M ⊙ yr-1, a stellar mass of {10}9.8+/- 0.3 M ⊙, and a half-light radius of {r}1/2=2.6+/- 0.7 {kpc}, typical of a main-sequence star-forming galaxy at z˜ 2. However, the Hα kinematics show a surprisingly tranquil <span class="hlt">velocity</span> <span class="hlt">field</span> with an ordered rotation ({V}{{c}}=200+/- 12 km s-1) and uniformly small <span class="hlt">velocity</span> dispersions ({V}σ ,{mean}=23 +/- 4 km s-1 and {V}σ ,{outer - {disk}}=15+/- 2 km s-1). The low gas <span class="hlt">velocity</span> dispersion is similar to local spiral galaxies and is consistent with the classic density wave theory where spiral arms form in dynamically cold and thin disks. We speculate that A1689B11 belongs to a population of rare spiral galaxies at z≳ 2 that mark the formation epoch of thin disks. Future <span class="hlt">observations</span> with the James Webb Space Telescope will greatly increase the sample of these rare galaxies and unveil the earliest onset of spiral arms.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015SPIE.9413E..35P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015SPIE.9413E..35P"><span>Image registration using stationary <span class="hlt">velocity</span> <span class="hlt">fields</span> parameterized by norm-minimizing Wendland kernel</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pai, Akshay; Sommer, Stefan; Sørensen, Lauge; Darkner, Sune; Sporring, Jon; Nielsen, Mads</p> <p>2015-03-01</p> <p>Interpolating kernels are crucial to solving a stationary <span class="hlt">velocity</span> <span class="hlt">field</span> (SVF) based image registration problem. This is because, <span class="hlt">velocity</span> <span class="hlt">fields</span> need to be computed in non-integer locations during integration. The regularity in the solution to the SVF registration problem is controlled by the regularization term. In a variational formulation, this term is traditionally expressed as a squared norm which is a scalar inner product of the interpolating kernels parameterizing the <span class="hlt">velocity</span> <span class="hlt">fields</span>. The minimization of this term using the standard spline interpolation kernels (linear or cubic) is only approximative because of the lack of a compatible norm. In this paper, we propose to replace such interpolants with a norm-minimizing interpolant - the Wendland kernel which has the same computational simplicity like B-Splines. An application on the Alzheimer's disease neuroimaging initiative showed that Wendland SVF based measures separate (Alzheimer's disease v/s normal controls) better than both B-Spline SVFs (p<0.05 in amygdala) and B-Spline freeform deformation (p<0.05 in amygdala and cortical gray matter).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5396373','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5396373"><span>Measurement of electroosmotic and electrophoretic <span class="hlt">velocities</span> using pulsed and sinusoidal electric <span class="hlt">fields</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>Sadek, Samir H.; Pimenta, Francisco; Pinho, Fernando T.</p> <p>2017-01-01</p> <p>In this work, we explore two methods to simultaneously measure the electroosmotic mobility in microchannels and the electrophoretic mobility of micron‐sized tracer particles. The first method is based on imposing a pulsed electric <span class="hlt">field</span>, which allows to isolate electrophoresis and electroosmosis at the startup and shutdown of the pulse, respectively. In the second method, a sinusoidal electric <span class="hlt">field</span> is generated and the mobilities are found by minimizing the difference between the measured <span class="hlt">velocity</span> of tracer particles and the <span class="hlt">velocity</span> computed from an analytical expression. Both methods produced consistent results using polydimethylsiloxane microchannels and polystyrene micro‐particles, provided that the temporal resolution of the particle tracking velocimetry technique used to compute the <span class="hlt">velocity</span> of the tracer particles is fast enough to resolve the diffusion time‐scale based on the characteristic channel length scale. Additionally, we present results with the pulse method for viscoelastic fluids, which show a more complex transient response with significant <span class="hlt">velocity</span> overshoots and undershoots after the start and the end of the applied electric pulse, respectively. PMID:27990654</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940031468','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940031468"><span>The <span class="hlt">velocity</span> <span class="hlt">field</span> created by a shallow bump in a boundary layer</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Gaster, Michael; Grosch, Chester E.; Jackson, Thomas L.</p> <p>1994-01-01</p> <p>We report the results of measurements of the disturbance <span class="hlt">velocity</span> <span class="hlt">field</span> generated in a boundary layer by a shallow three-dimensional bump oscillating at a very low frequency on the surface of a flat plate. Profiles of the mean <span class="hlt">velocity</span>, the disturbance <span class="hlt">velocity</span> at the fundamental frequency and at the first harmonic are presented. These profiles were measured both upstream and downstream of the oscillating bump. Measurements of the disturbance <span class="hlt">velocity</span> were also made at various spanwise and downstream locations at a fixed distance from the boundary of one displacement thickness. Finally, the spanwise spectrum of the disturbances at three locations downstream of the bump are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015MNRAS.453L.108L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015MNRAS.453L.108L"><span>Filaments from the galaxy distribution and from the <span class="hlt">velocity</span> <span class="hlt">field</span> 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>Libeskind, Noam I.; Tempel, Elmo; Hoffman, Yehuda; Tully, R. Brent; Courtois, Hélène</p> <p>2015-10-01</p> <p>The cosmic web that characterizes the large-scale structure of the Universe can be quantified by a variety of methods. For example, large redshift surveys can be used in combination with point process algorithms to extract long curvilinear filaments in the galaxy distribution. Alternatively, given a full 3D reconstruction of the <span class="hlt">velocity</span> <span class="hlt">field</span>, kinematic techniques can be used to decompose the web into voids, sheets, filaments and knots. In this Letter, we look at how two such algorithms - the Bisous model and the <span class="hlt">velocity</span> shear web - compare with each other in the local Universe (within 100 Mpc), finding good agreement. This is both remarkable and comforting, given that the two methods are radically different in ideology and applied to completely independent and different data sets. Unsurprisingly, the methods are in better agreement when applied to unbiased and complete data sets, like cosmological simulations, than when applied to <span class="hlt">observational</span> samples. We conclude that more <span class="hlt">observational</span> data is needed to improve on these methods, but that both methods are most likely properly tracing the underlying distribution of matter in the Universe.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70029582','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70029582"><span>Transient rheology of the upper mantle beneath central Alaska inferred from the crustal <span class="hlt">velocity</span> <span class="hlt">field</span> following the 2002 Denali earthquake</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>2005-01-01</p> <p>The M7.9 2002 Denali earthquake, Alaska, is one of the largest strike-slip earthquakes ever recorded. The postseismic GPS <span class="hlt">velocity</span> <span class="hlt">field</span> around the 300-km-long rupture is characterized by very rapid horizontal <span class="hlt">velocity</span> up to ???300 mm/yr for the first 0.1 years and slower but still elevated horizontal <span class="hlt">velocity</span> up to ???100 mm/yr for the succeeding 1.5 years. I find that the spatial and temporal pattern of the displacement <span class="hlt">field</span> may be explained by a transient mantle rheology. Representing the regional upper mantle as a Burghers body, I infer steady state and transient viscosities of ??1 = 2.8 ?? 1018 Pa s and ??2 = 1.0 ?? 1017 Pa s, respectively, corresponding to material relaxation times of 1.3 and 0.05 years. The lower crustal viscosity is poorly constrained by the considered horizontal <span class="hlt">velocity</span> <span class="hlt">field</span>, and the quoted mantle viscosities assume a steady state lower crust viscosity that is 7??1. Systematic bias in predicted versus <span class="hlt">observed</span> <span class="hlt">velocity</span> vectors with respect to a fixed North America during the first 3-6 months following the earthquake is reduced when all <span class="hlt">velocity</span> vectors are referred to a fixed site. This suggests that the post-Denali GPS time series for the first 1.63 years are shaped by a combination of a common mode noise source during the first 3-6 months plus viscoelastic relaxation controlled by a transient mantle rheology.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19850036090&hterms=IOTA&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DIOTA','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19850036090&hterms=IOTA&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DIOTA"><span>Copernicus <span class="hlt">observations</span> of Iota Herculis <span class="hlt">velocity</span> variations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rogerson, J. B., Jr.</p> <p>1984-01-01</p> <p><span class="hlt">Observations</span> 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-<span class="hlt">velocity</span> 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.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10329E..3PG','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10329E..3PG"><span>An optical flow-based method for <span class="hlt">velocity</span> <span class="hlt">field</span> of fluid flow estimation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Głomb, Grzegorz; Świrniak, Grzegorz; Mroczka, Janusz</p> <p>2017-06-01</p> <p>The aim of this paper is to present a method for estimating flow-<span class="hlt">velocity</span> vector <span class="hlt">fields</span> using the Lucas-Kanade algorithm. The optical flow measurements are based on the Particle Image Velocimetry (PIV) technique, which is commonly used in fluid mechanics laboratories in both research institutes and industry. Common approaches for an optical characterization of <span class="hlt">velocity</span> <span class="hlt">fields</span> base on computation of partial derivatives of the image intensity using finite differences. Nevertheless, the accuracy of <span class="hlt">velocity</span> <span class="hlt">field</span> computations is low due to the fact that an exact estimation of spatial derivatives is very difficult in presence of rapid intensity changes in the PIV images, caused by particles having small diameters. The method discussed in this paper solves this problem by interpolating the PIV images using Gaussian radial basis functions. This provides a significant improvement in the accuracy of the <span class="hlt">velocity</span> estimation but, more importantly, allows for the evaluation of the derivatives in intermediate points between pixels. Numerical analysis proves that the method is able to estimate even a separate vector for each particle with a 5× 5 px2 window, whereas a classical correlation-based method needs at least 4 particle images. With the use of a specialized multi-step hybrid approach to data analysis the method improves the estimation of the particle displacement far above 1 px.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DPPJ11076H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DPPJ11076H"><span>Direct comparison of neutral <span class="hlt">velocity</span> distribution measurements and simulations in the vicinity of an absorbing boundary oblique to a magnetic <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Henriquez, Miguel F.; Thompson, Derek S.; Keniley, Shane; Curreli, Davide; Steinberger, Thomas E.; Caron, David D.; Jemiolo, Andrew J.; McLaughlin, Jacob W.; Dufor, Mikal T.; Neal, Luke A.; Scime, Earl E.; Siddiqui, M. Umair</p> <p>2017-10-01</p> <p>Plasma-boundary interactions are strongly affected by the sheath and presheath structures that form near the boundary surface. Recent measurements have <span class="hlt">observed</span> ion transport across magnetic <span class="hlt">field</span> lines in regions where the surface is oblique to the background magnetic <span class="hlt">field</span> (ψ =74°) . In these boundary regions, charge exchange collisions may provide a mechanism through which neutral particles interact with the long distance presheath electric <span class="hlt">field</span>. We report efforts to directly compare Boltzmann and particle-in-cell simulations with 3D neutral <span class="hlt">velocity</span> distribution functions (NVDFs) using laser induced fluorescence (LIF) in a magnetized plasma boundary region. We present a novel LIF method for measuring Ar-II metastable <span class="hlt">velocity</span> distributions, in which we <span class="hlt">observe</span> the 738.6014 nm fluorescence (2p3 to 1s4 in Paschen's notation), that results from absorption of the 706.9167 nm (1s5 metastable to 2p3) pump laser, providing neutral temperatures and flows. We additionally describe electrostatic probe measurements in the same region.</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/2009JGRA..114.0E05K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009JGRA..114.0E05K"><span>An explanation for parallel electric <span class="hlt">field</span> pulses <span class="hlt">observed</span> over thunderstorms</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kelley, M. C.; Barnum, B. H.</p> <p>2009-10-01</p> <p>Every electric <span class="hlt">field</span> instrument flown on sounding rockets over a thunderstorm has detected pulses of electric <span class="hlt">fields</span> parallel to the Earth's magnetic <span class="hlt">field</span> associated with every strike. This paper describes the ionospheric signatures found during a flight from Wallops Island, Virginia, on 2 September 1995. The electric <span class="hlt">field</span> results in a drifting Maxwellian corresponding to energies up to 1 eV. The distribution function relaxes because of elastic and inelastic collisions, resulting in electron heating up to 4000-5000 K and potentially <span class="hlt">observable</span> red line emissions and enhanced ISR electron temperatures. The <span class="hlt">field</span> strength scales with the current in cloud-to-ground strikes and falls off as r -1 with distance. Pulses of both polarities are found, although most electric <span class="hlt">fields</span> are downward, parallel to the magnetic <span class="hlt">field</span>. The pulse may be the reaction of ambient plasma to a current pulse carried at the whistler packet's highest group <span class="hlt">velocity</span>. The charge source required to produce the electric <span class="hlt">field</span> is very likely electrons of a few keV traveling at the packet <span class="hlt">velocity</span>. We conjecture that the current source is the divergence of the current flowing at mesospheric heights, the phenomenon called an elve. The whistler packet's effective radiated power is as high as 25 mW at ionospheric heights, comparable to some ionospheric heater transmissions. Comparing the Poynting flux at the base of the ionosphere with flux an equal distance away along the ground, some 30 db are lost in the mesosphere. Another 10 db are lost in the transition from free space to the whistler mode.</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> <span class="hlt">fields</span>. 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('https://ntrs.nasa.gov/search.jsp?R=19750044668&hterms=lazarus&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D80%26Ntt%3Dlazarus','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19750044668&hterms=lazarus&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D80%26Ntt%3Dlazarus"><span>Pioneer 7 <span class="hlt">observations</span> of plasma flow and <span class="hlt">field</span> reversal regions in the distant geomagnetic tail</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Walker, R. C.; Lazarus, A. J.; Villante, U.</p> <p>1975-01-01</p> <p>The present paper gives the results of an extensive analysis of plasma and magnetic-<span class="hlt">field</span> data from Pioneer 7 taken in the geomagnetic tail approximately 1000 earth radii downstream from earth. The principal <span class="hlt">observations</span> are: (1) measurable fluxes of protons in the tail, flowing away from earth, sometimes with a double-peaked <span class="hlt">velocity</span> distribution; (2) <span class="hlt">field</span> reversal regions in which the <span class="hlt">field</span> changes from radial to antiradial by a vector rotation in the north-south plane; and (3) general characteristics of the tail similar to those <span class="hlt">observed</span> near earth with good correlation between taillike magnetic <span class="hlt">fields</span> and plasma.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012EGUGA..1413912F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012EGUGA..1413912F"><span>Multifractal Analysis of <span class="hlt">Velocity</span> Vector <span class="hlt">Fields</span> and a Continuous In-Scale Cascade Model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fitton, G.; Tchiguirinskaia, I.; Schertzer, D.; Lovejoy, S.</p> <p>2012-04-01</p> <p>In this study we have compared the multifractal analyses of small-scale surface-layer wind <span class="hlt">velocities</span> from two different datasets. The first dataset consists of six-months of wind <span class="hlt">velocity</span> and temperature measurements at the heights 22, 23 and 43m. The measurements came from 3D sonic anemometers with a 10Hz data output rate positioned on a mast in a wind farm test site subject to wake turbulence effects. The location of the test site (Corsica, France) meant the large scale structures were subject to topography effects that therefore possibly caused buoyancy effects. The second dataset (Germany) consists of 300 twenty minute samples of horizontal wind <span class="hlt">velocity</span> magnitudes simultaneously recorded at several positions on two masts. There are eight propeller anemometers on each mast, recording <span class="hlt">velocity</span> magnitude data at 2.5Hz. The positioning of the anemometers is such that there are effectively two grids. One grid of 3 rows by 4 columns and a second of 5 rows by 2 columns. The ranges of temporal scale over which the analyses were done were from 1 to 103 seconds for both datasets. Thus, under the universal multifractal framework we found both datasets exhibit parameters α ≈ 1.5 and C1 ≈ 0.1. The parameters α and C1, measure respectively the multifractality and mean intermittency of the scaling <span class="hlt">field</span>. A third parameter, H, quantifies the divergence from conservation of the <span class="hlt">field</span> (e.g. H = 0 for the turbulent energy flux density). To estimate the parameters we used the ratio of the scaling moment function of the energy flux and of the <span class="hlt">velocity</span> increments. This method was particularly useful when estimating the parameter α over larger scales. In fact it was not possible to obtain a reasonable estimate of alpha using the usual double trace moment method. For each case the scaling behaviour of the wind was almost isotropic when the scale ranges remained close to the sphero-scale. For the Corsica dataset this could be seen by the agreement of the spectral exponents of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhRvL.112v1102H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhRvL.112v1102H"><span>Clear and Measurable Signature of Modified Gravity in the Galaxy <span class="hlt">Velocity</span> <span class="hlt">Field</span></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.; Barreira, Alexandre; Frenk, Carlos S.; Li, Baojiu; Cole, Shaun</p> <p>2014-06-01</p> <p>The <span class="hlt">velocity</span> <span class="hlt">field</span> of dark matter and galaxies reflects the continued action of gravity throughout cosmic history. We show that the low-order moments of the pairwise <span class="hlt">velocity</span> distribution v12 are a powerful diagnostic of the laws of gravity on cosmological scales. In particular, the projected line-of-sight galaxy pairwise <span class="hlt">velocity</span> dispersion σ12(r) is very sensitive to the presence of modified gravity. Using a set of high-resolution N-body simulations, we compute the pairwise <span class="hlt">velocity</span> distribution and its projected line-of-sight dispersion for a class of modified gravity theories: the chameleon f(R) gravity and Galileon gravity (cubic and quartic). The <span class="hlt">velocities</span> of dark matter halos with a wide range of masses would exhibit deviations from general relativity at the (5-10)σ level. We examine strategies for detecting these deviations in galaxy redshift and peculiar <span class="hlt">velocity</span> surveys. If detected, this signature would be a "smoking gun" for modified gravity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016MeScT..27i4010E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016MeScT..27i4010E"><span>A hybrid experimental-numerical technique for determining 3D <span class="hlt">velocity</span> <span class="hlt">fields</span> from planar 2D PIV data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Eden, A.; Sigurdson, M.; Mezić, I.; Meinhart, C. D.</p> <p>2016-09-01</p> <p>Knowledge of 3D, three component <span class="hlt">velocity</span> <span class="hlt">fields</span> is central to the understanding and development of effective microfluidic devices for lab-on-chip mixing applications. In this paper we present a hybrid experimental-numerical method for the generation of 3D flow information from 2D particle image velocimetry (PIV) experimental data and finite element simulations of an alternating current electrothermal (ACET) micromixer. A numerical least-squares optimization algorithm is applied to a theory-based 3D multiphysics simulation in conjunction with 2D PIV data to generate an improved estimation of the steady state <span class="hlt">velocity</span> <span class="hlt">field</span>. This 3D <span class="hlt">velocity</span> <span class="hlt">field</span> can be used to assess mixing phenomena more accurately than would be possible through simulation alone. Our technique can also be used to estimate uncertain quantities in experimental situations by fitting the gathered <span class="hlt">field</span> data to a simulated physical model. The optimization algorithm reduced the root-mean-squared difference between the experimental and simulated <span class="hlt">velocity</span> <span class="hlt">fields</span> in the target region by more than a factor of 4, resulting in an average error less than 12% of the average <span class="hlt">velocity</span> magnitude.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018CNSNS..59..553G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018CNSNS..59..553G"><span>Mean-<span class="hlt">field</span> <span class="hlt">velocity</span> difference model considering the average effect of multi-vehicle interaction</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Guo, Yan; Xue, Yu; Shi, Yin; Wei, Fang-ping; Lü, Liang-zhong; He, Hong-di</p> <p>2018-06-01</p> <p>In this paper, a mean-<span class="hlt">field</span> <span class="hlt">velocity</span> difference model(MFVD) is proposed to describe the average effect of multi-vehicle interactions on the whole road. By stability analysis, the stability condition of traffic system is obtained. Comparison with stability of full <span class="hlt">velocity</span>-difference (FVD) model and the completeness of MFVD model are discussed. The mKdV equation is derived from MFVD model through nonlinear analysis to reveal the traffic jams in the form of the kink-antikink density wave. Then the numerical simulation is performed and the results illustrate that the average effect of multi-vehicle interactions plays an important role in effectively suppressing traffic jam. The increase strength of the mean-<span class="hlt">field</span> <span class="hlt">velocity</span> difference in MFVD model can rapidly reduce traffic jam and enhance the stability of traffic system.</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 <span class="hlt">field</span> 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/2017GeoJI.208..246H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GeoJI.208..246H"><span>Evaluating a campaign GNSS <span class="hlt">velocity</span> <span class="hlt">field</span> derived from an online precise point positioning service</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Holden, L.; Silcock, D.; Choy, S.; Cas, R.; Ailleres, L.; Fournier, N.</p> <p>2017-01-01</p> <p>Traditional processing of Global Navigation Satellite System (GNSS) data using dedicated scientific software has provided the highest levels of positional accuracy, and has been used extensively in geophysical deformation studies. To achieve these accuracies a significant level of understanding and training is required, limiting their availability to the general scientific community. Various online GNSS processing services, now freely available, address some of these difficulties and allow users to easily process their own GNSS data and potentially obtain high quality results. Previous research into these services has focused on Continually Operating Reference Station (CORS) GNSS data. Less research exists on the results achievable with these services using large campaign GNSS data sets, which are inherently noisier than CORS data. Even less research exists on the quality of <span class="hlt">velocity</span> <span class="hlt">fields</span> derived from campaign GNSS data processed through online precise point positioning services. Particularly, whether they are suitable for geodynamic and deformation studies where precise and reliable <span class="hlt">velocities</span> are needed. In this research, we process a very large campaign GPS data set (spanning 10 yr) with the online Jet Propulsion Laboratory Automated Precise Positioning Service. This data set is taken from a GNSS network specifically designed and surveyed to measure deformation through the central North Island of New Zealand. This includes regional CORS stations. We then use these coordinates to derive a horizontal and vertical <span class="hlt">velocity</span> <span class="hlt">field</span>. This is the first time that a large campaign GPS data set has been processed solely using an online service and the solutions used to determine a horizontal and vertical <span class="hlt">velocity</span> <span class="hlt">field</span>. We compared this <span class="hlt">velocity</span> <span class="hlt">field</span> to that of another well utilized GNSS scientific software package. The results show a good agreement between the CORS positions and campaign station <span class="hlt">velocities</span> obtained from the two approaches. We discuss the implications</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015TESS....130505D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015TESS....130505D"><span>MAVEN <span class="hlt">Observations</span> of Escaping Planetary Ions from the Martian Atmosphere: Mass, <span class="hlt">Velocity</span>, and Spatial Distributions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dong, Yaxue; Fang, Xiaohua; Brain, D. A.; McFadden, James P.; Halekas, Jasper; Connerney, Jack</p> <p>2015-04-01</p> <p>The Mars-solar wind interaction accelerates and transports planetary ions away from the Martian atmosphere through a number of processes, including ‘pick-up’ by electromagnetic <span class="hlt">fields</span>. The MAVEN spacecraft has made routine <span class="hlt">observations</span> of escaping planetary ions since its arrival at Mars in September 2014. The SupraThermal And Thermal Ion Composition (STATIC) instrument measures the ion energy, mass, and angular spectra. It has detected energetic planetary ions during most of the spacecraft orbits, which are attributed to the pick-up process. We found significant variations in the escaping ion mass and <span class="hlt">velocity</span> distributions from the STATIC data, which can be explained by factors such as varying solar wind conditions, contributions of particles from different source locations and different phases during the pick-up process. We also study the spatial distributions of different planetary ion species, which can provide insight into the physics of ion escaping process and enhance our understanding of atmospheric erosion by the solar wind. Our results will be further interpreted within the context of the upstream solar wind conditions measured by the MAVEN Solar Wind Ion Analyzer (SWIA) instrument and the magnetic <span class="hlt">field</span> environment measured by the Magnetometer (MAG) instrument. Our study shows that the ion spatial distribution in the Mars-Sun-Electric-<span class="hlt">Field</span> (MSE) coordinate system and the <span class="hlt">velocity</span> space distribution with respect to the local magnetic <span class="hlt">field</span> line can be used to distinguish the ions escaping through the polar plume and those through the tail region. The contribution of the polar plume ion escape to the total escape rate will also be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1912378D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1912378D"><span>Dilution and Mixing in transient <span class="hlt">velocity</span> <span class="hlt">fields</span>: a first-order analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Di Dato, Mariaines; de Barros, Felipe, P. J.; Fiori, Aldo; Bellin, Alberto</p> <p>2017-04-01</p> <p>An appealing remediation technique is in situ oxidation, which effectiveness is hampered by difficulties in obtaining good mixing of the injected oxidant with the contaminant, particularly when the contaminant plume is contained and therefore its deformation is physically constrained. Under such conditions (i.e. containment), mixing may be augmented by inducing temporal fluctuations of the <span class="hlt">velocity</span> <span class="hlt">field</span>. The temporal variability of the flow <span class="hlt">field</span> may increase the deformation of the plume such that diffusive mass flux becomes more effective. A transient periodic <span class="hlt">velocity</span> <span class="hlt">field</span> can be obtained by an engineered sequence of injections and extractions from wells, which may serve also as a hydraulic barrier to confine the plume. Assessing the effectiveness of periodic flows to maximize solute mixing is a difficult task given the need to use a 3D setup and the large number of possible flow configurations that should be analyzed in order to identify the optimal one. This is the typical situation in which analytical solutions, though approximated, may assist modelers in screening possible alternative flow configurations such that solute dilution is maximized. To quantify dilution (i.e. a precondition that enables reactive mixing) we utilize the concept of the dilution index [1]. In this presentation, the periodic flow takes place in an aquifer with spatially variable hydraulic conductivity <span class="hlt">field</span> which is modeled as a Stationary Spatial Random Function. We developed a novel first-order analytical solution of the dilution index under the hypothesis that the flow can be approximated as a sequence of steady state configurations with the mean <span class="hlt">velocity</span> changing with time in intensity and direction. This is equivalent to assume that the characteristic time of the transient behavior is small compared to the period characterizing the change in time of the mean <span class="hlt">velocity</span>. A few closed paths have been analyzed quantifying their effectiveness in enhancing dilution and thereby mixing</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006QJRMS.132..125L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006QJRMS.132..125L"><span>Estimating vertical <span class="hlt">velocity</span> and radial flow from Doppler radar <span class="hlt">observations</span> of tropical cyclones</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lee, J. L.; Lee, W. C.; MacDonald, A. E.</p> <p>2006-01-01</p> <p>The mesoscale vorticity method (MVM) is used in conjunction with the ground-based <span class="hlt">velocity</span> track display (GBVTD) to derive the inner-core vertical <span class="hlt">velocity</span> from Doppler radar <span class="hlt">observations</span> of tropical cyclone (TC) Danny (1997). MVM derives the vertical <span class="hlt">velocity</span> 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 <span class="hlt">velocity</span> 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 <span class="hlt">observations</span> 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 <span class="hlt">velocity</span> 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 <span class="hlt">observed</span> by Doppler radars to derive realistic vertical <span class="hlt">velocity</span> 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 <span class="hlt">observing</span> system such as state-of-the-art Doppler radars.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24675957','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24675957"><span><span class="hlt">Observation</span> of Brownian motion in liquids at short times: instantaneous <span class="hlt">velocity</span> and memory loss.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kheifets, Simon; Simha, Akarsh; Melin, Kevin; Li, Tongcang; Raizen, Mark G</p> <p>2014-03-28</p> <p>Measurement of the instantaneous <span class="hlt">velocity</span> of Brownian motion of suspended particles in liquid probes the microscopic foundations of statistical mechanics in soft condensed matter. However, instantaneous <span class="hlt">velocity</span> has eluded experimental <span class="hlt">observation</span> 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 <span class="hlt">observe</span> the hydrodynamic instantaneous <span class="hlt">velocity</span> 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 <span class="hlt">observe</span> an anticorrelated thermal force, which is conventionally assumed to be uncorrelated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19820054845&hterms=Magnetic+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DMagnetic%2Benergy','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19820054845&hterms=Magnetic+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DMagnetic%2Benergy"><span>Vector magnetic <span class="hlt">field</span> evolution, energy storage, and associated photospheric <span class="hlt">velocity</span> shear within a flare-productive active region</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Krall, K. R.; Smith, J. B., Jr.; Hagyard, M. J.; West, E. A.; Cummings, N. P.</p> <p>1982-01-01</p> <p>Sheared photospheric <span class="hlt">velocity</span> <span class="hlt">fields</span> inferred from spot motions for April 5-7, 1980, are compared with both transverse magnetic <span class="hlt">field</span> orientation changes and with the region's flare history. Rapid spot motions and high inferred <span class="hlt">velocity</span> shear coincide with increased <span class="hlt">field</span> alignment along the longitudinal neutral line and with increased flare activity, while a later decrease in <span class="hlt">velocity</span> shear precedes a more relaxed magnetic configuration and decrease in flare activity. It is estimated that magnetic reconfiguration produced by the relative <span class="hlt">velocities</span> of the spots could cause storage of about 10 to the 32nd erg/day, while flares occurring during this time expended no more than about 10 to the 31st erg/day.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010ASSP...19..517M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010ASSP...19..517M"><span>Magnetic and <span class="hlt">Velocity</span> <span class="hlt">Field</span> Changes Related to the Solar Flares of 28 and 29 October 2003</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Maurya, R. A.; Ambastha, A.</p> <p></p> <p>Magnetic and <span class="hlt">velocity</span> <span class="hlt">field</span> measurements of solar active regions suffer from ambiguities caused by the change in spectral line profiles that occur during the impulsive phase of a major flare. This leads to difficulties in correct interpretation of any flare-related changes. Using magnetic and Doppler movies taken with GONG and MDI, we have detected transient, "moving" features around the peak phases of the X17.2/4B flare <span class="hlt">observed</span> on 28 October 2003 and the X10/2B flare <span class="hlt">observed</span> on 29 October 2003 in super-active region NOAA 10486. These features were located near the compact acoustic sources reported earlier by Donea and Lindsey (2005) and the seismic sources reported by Zharkova and Zharkov (2007).We find a moving feature, spatially and temporally associated with the flare ribbons, that separates away at speeds ranging from 30 to 50 km s-1 as <span class="hlt">observed</span> in photospheric white light and in temperature-minimum (1600 Å), chromospheric (Hα), and transition-region (284Å ) intensities.We suggest that such moving features arise from the line-profile changes attributed to downward electron jets associated with the flare, and do not reflect real changes in the photospheric magnetic and <span class="hlt">velocity</span> <span class="hlt">fields</span>. However, abrupt and persistent changes in the pre- and post-flare phases were also found, which do not seem to be affected by line-profile changes. The detailed results have been appeared in Maurya and Ambastha (2009).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003AGUFMPP31C0273B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003AGUFMPP31C0273B"><span>Antarctic Glaciological Data at NSIDC: <span class="hlt">field</span> data, temperature, and ice <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>Bauer, R.; Bohlander, J.; Scambos, T.; Berthier, E.; Raup, B.; Scharfen, G.</p> <p>2003-12-01</p> <p>An extensive collection of many Antarctic glaciological parameters is available for the polar science community upon request. The National Science Foundation's Office of Polar Programs funds the Antarctic Glaciological Data Center (AGDC) at the National Snow and Ice Data Center (NSIDC) to archive and distribute Antarctic glaciological and cryospheric system data collected by the U.S. Antarctic Program. AGDC facilitates data exchange among Principal Investigators, preserves recently collected data useful to future research, gathers data sets from past research, and compiles continent-wide information useful for modeling and <span class="hlt">field</span> work planning. Data sets are available via our web site, http://nsidc.org/agdc/. From here, users can access extensive documentation, citation information, locator maps, derived images and references, and the numerical data. More than 50 Antarctic scientists have contributed data to the archive. Among the compiled products distributed by AGDC are VELMAP and THERMAP. THERMAP is a compilation of over 600 shallow firn temperature measurements ('10-meter temperatures') collected since 1950. These data provide a record of mean annual temperature, and potentially hold a record of climate change on the continent. The data are represented with maps showing the traverse route, and include data sources, measurement technique, and additional measurements made at each site, i.e., snow density and accumulation. VELMAP is an archive of surface ice <span class="hlt">velocity</span> measurements for the Antarctic Ice Sheet. The primary objective of VELMAP is to assemble a historic record of outlet glaciers and ice shelf ice motion over the Antarctic. The collection includes both PI-contributed measurements and data generated at NSIDC using Landsat and SPOT satellite imagery. Tabular data contain position, speed, bearing, and data quality information, and related references. Two new VELMAP data sets are highlighted: the Mertz Glacier and the Institute Ice Stream. Mertz Glacier ice</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 <span class="hlt">field</span> 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 <span class="hlt">field</span> 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 <span class="hlt">field</span>. 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/2015PhFl...27b5102P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhFl...27b5102P"><span>Statistical analysis of the <span class="hlt">velocity</span> and scalar <span class="hlt">fields</span> in reacting turbulent wall-jets</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pouransari, Z.; Biferale, L.; Johansson, A. V.</p> <p>2015-02-01</p> <p>The concept of local isotropy in a chemically reacting turbulent wall-jet flow is addressed using direct numerical simulation (DNS) data. Different DNS databases with isothermal and exothermic reactions are examined. The chemical reaction and heat release effects on the turbulent <span class="hlt">velocity</span>, passive scalar, and reactive species <span class="hlt">fields</span> are studied using their probability density functions (PDFs) and higher order moments for <span class="hlt">velocities</span> and scalar <span class="hlt">fields</span>, as well as their gradients. With the aid of the anisotropy invariant maps for the Reynolds stress tensor, the heat release effects on the anisotropy level at different wall-normal locations are evaluated and found to be most accentuated in the near-wall region. It is <span class="hlt">observed</span> that the small-scale anisotropies are persistent both in the near-wall region and inside the jet flame. Two exothermic cases with different Damköhler numbers are examined and the comparison revealed that the Damköhler number effects are most dominant in the near-wall region, where the wall cooling effects are influential. In addition, with the aid of PDFs conditioned on the mixture fraction, the significance of the reactive scalar characteristics in the reaction zone is illustrated. We argue that the combined effects of strong intermittency and strong persistency of anisotropy at the small scales in the entire domain can affect mixing and ultimately the combustion characteristics of the reacting flow.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMEP52B..03T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMEP52B..03T"><span>Does resolution of flow <span class="hlt">field</span> <span class="hlt">observation</span> influence apparent habitat use and energy expenditure in juvenile coho salmon?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tullos, D. D.; Walter, C.; Dunham, J.</p> <p>2016-12-01</p> <p>This study investigated how the resolution of <span class="hlt">observation</span> influences interpretation of how fish, juvenile Coho Salmon (Oncorhynchus kisutch), exploit the hydraulic environment in streams. Our objectives were to evaluate how spatial resolution of the flow <span class="hlt">field</span> <span class="hlt">observation</span> influenced: 1) the <span class="hlt">velocities</span> considered to be representative of habitat units; 2) patterns of use of the hydraulic environment by fish; and 3) estimates of energy expenditure. We addressed these objectives using <span class="hlt">observations</span> within a 1:1 scale physical model of a full-channel log jam in an outdoor experimental stream. <span class="hlt">Velocities</span> were measured with Acoustic Doppler Velocimetry at a 10 cm grid spacing, whereas fish locations and tailbeat frequencies were documented over time using underwater videogrammetry. Results highlighted that resolution of <span class="hlt">observation</span> did impact perceived habitat use and energy expenditure, as did the location of measurement within habitat units and the use of averaging to summarize <span class="hlt">velocities</span> within a habitat unit. In this experiment, the range of <span class="hlt">velocities</span> and energy expenditure estimates increased with coarsening resolution, reducing the likelihood of measuring the <span class="hlt">velocities</span> locally experienced by fish. In addition, the coarser resolutions contributed to fish appearing to select <span class="hlt">velocities</span> that were higher than what was measured at finer resolutions. These findings indicate the need for careful attention to and communication of resolution of <span class="hlt">observation</span> in investigating the hydraulic environment and in determining the habitat needs and bioenergetics of aquatic biota.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880056860&hterms=environnement&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Denvironnement','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880056860&hterms=environnement&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Denvironnement"><span>Using a constraint on the parallel <span class="hlt">velocity</span> when determining electric <span class="hlt">fields</span> with EISCAT</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Caudal, G.; Blanc, M.</p> <p>1988-01-01</p> <p>A method is proposed to determine the perpendicular components of the ion <span class="hlt">velocity</span> vector (and hence the perpendicular electric <span class="hlt">field</span>) from EISCAT tristatic measurements, in which one introduces an additional constraint on the parallel <span class="hlt">velocity</span>, in order to take account of our knowledge that the parallel <span class="hlt">velocity</span> of ions is small. This procedure removes some artificial features introduced when the tristatic geometry becomes too unfavorable. It is particularly well suited for the southernmost or northernmost positions of the tristatic measurements performed by meridian scan experiments (CP3 mode).</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/2018MNRAS.473.3454B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.473.3454B"><span>Filament formation in wind-cloud interactions- II. Clouds with turbulent density, <span class="hlt">velocity</span>, and magnetic <span class="hlt">fields</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Banda-Barragán, W. E.; Federrath, C.; Crocker, R. M.; Bicknell, G. V.</p> <p>2018-01-01</p> <p>We present a set of numerical experiments designed to systematically investigate how turbulence and magnetic <span class="hlt">fields</span> influence the morphology, energetics, and dynamics of filaments produced in wind-cloud interactions. We cover 3D, magnetohydrodynamic systems of supersonic winds impacting clouds with turbulent density, <span class="hlt">velocity</span>, and magnetic <span class="hlt">fields</span>. We find that lognormal density distributions aid shock propagation through clouds, increasing their <span class="hlt">velocity</span> dispersion and producing filaments with expanded cross-sections and highly magnetized knots and subfilaments. In self-consistently turbulent scenarios, the ratio of filament to initial cloud magnetic energy densities is ∼1. The effect of Gaussian <span class="hlt">velocity</span> <span class="hlt">fields</span> is bound to the turbulence Mach number: Supersonic <span class="hlt">velocities</span> trigger a rapid cloud expansion; subsonic <span class="hlt">velocities</span> only have a minor impact. The role of turbulent magnetic <span class="hlt">fields</span> depends on their tension and is similar to the effect of radiative losses: the stronger the magnetic <span class="hlt">field</span> or the softer the gas equation of state, the greater the magnetic shielding at wind-filament interfaces and the suppression of Kelvin-Helmholtz instabilities. Overall, we show that including turbulence and magnetic <span class="hlt">fields</span> is crucial to understanding cold gas entrainment in multiphase winds. While cloud porosity and supersonic turbulence enhance the acceleration of clouds, magnetic shielding protects them from ablation and causes Rayleigh-Taylor-driven subfilamentation. Wind-swept clouds in turbulent models reach distances ∼15-20 times their core radius and acquire bulk speeds ∼0.3-0.4 of the wind speed in one cloud-crushing time, which are three times larger than in non-turbulent models. In all simulations, the ratio of turbulent magnetic to kinetic energy densities asymptotes at ∼0.1-0.4, and convergence of all relevant dynamical properties requires at least 64 cells per cloud radius.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10943050','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10943050"><span>Estimation of 3-D conduction <span class="hlt">velocity</span> vector <span class="hlt">fields</span> from cardiac mapping data.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Barnette, A R; Bayly, P V; Zhang, S; Walcott, G P; Ideker, R E; Smith, W M</p> <p>2000-08-01</p> <p>A method to estimate three-dimensional (3-D) conduction <span class="hlt">velocity</span> vector <span class="hlt">fields</span> in cardiac tissue is presented. The speed and direction of propagation are found from polynomial "surfaces" fitted to space-time (x, y, z, t) coordinates of cardiac activity. The technique is applied to sinus rhythm and paced rhythm mapped with plunge needles at 396-466 sites in the canine myocardium. The method was validated on simulated 3-D plane and spherical waves. For simulated data, conduction <span class="hlt">velocities</span> were estimated with an accuracy of 1%-2%. In experimental data, estimates of conduction speeds during paced rhythm were slower than those found during normal sinus rhythm. Vector directions were also found to differ between different types of beats. The technique was able to distinguish between premature ventricular contractions and sinus beats and between sinus and paced beats. The proposed approach to computing <span class="hlt">velocity</span> vector <span class="hlt">fields</span> provides an automated, physiological, and quantitative description of local electrical activity in 3-D tissue. This method may provide insight into abnormal conduction associated with fatal ventricular arrhythmias.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017A%26A...597A..10F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017A%26A...597A..10F"><span>A test <span class="hlt">field</span> for Gaia. Radial <span class="hlt">velocity</span> catalogue of stars in the South Ecliptic Pole</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Frémat, Y.; Altmann, M.; Pancino, E.; Soubiran, C.; Jofré, P.; Damerdji, Y.; Heiter, U.; Royer, F.; Seabroke, G.; Sordo, R.; Blanco-Cuaresma, S.; Jasniewicz, G.; Martayan, C.; Thévenin, F.; Vallenari, A.; Blomme, R.; David, M.; Gosset, E.; Katz, D.; Viala, Y.; Boudreault, S.; Cantat-Gaudin, T.; Lobel, A.; Meisenheimer, K.; Nordlander, T.; Raskin, G.; Royer, P.; Zorec, J.</p> <p>2017-01-01</p> <p>Context. Gaia is a space mission that is currently measuring the five astrometric parameters, as well as spectrophotometry of at least 1 billion stars to G = 20.7 mag with unprecedented precision. The sixth parameter in phase space (I.e., radial <span class="hlt">velocity</span>) is also measured thanks to medium-resolution spectroscopy that is being obtained for the 150 million brightest stars. During the commissioning phase, two <span class="hlt">fields</span>, one around each ecliptic pole, have been repeatedly <span class="hlt">observed</span> to assess and to improve the overall satellite performances, as well as the associated reduction and analysis software. A ground-based photometric and spectroscopic survey was therefore initiated in 2007, and is still running to gather as much information as possible about the stars in these <span class="hlt">fields</span>. This work is of particular interest to the validation of the radial <span class="hlt">velocity</span> spectrometer outputs. Aims: The paper presents the radial <span class="hlt">velocity</span> measurements performed for the Southern targets in the 12-17 R magnitude range on high- to mid-resolution spectra obtained with the GIRAFFE and UVES spectrographs. Methods: Comparison of the South Ecliptic Pole (SEP) GIRAFFE data to spectroscopic templates <span class="hlt">observed</span> with the HERMES (Mercator in La Palma, Spain) spectrograph enabled a first coarse characterisation of the 747 SEP targets. Radial <span class="hlt">velocities</span> were then obtained by comparing the results of three different methods. Results: In this paper, we present an initial overview of the targets to be found in the 1 sq. deg SEP region that was <span class="hlt">observed</span> repeatedly by Gaia ever since its commissioning. In our representative sample, we identified one galaxy, six LMC S-stars, nine candidate chromospherically active stars, and confirmed the status of 18 LMC Carbon stars. A careful study of the 3471 epoch radial <span class="hlt">velocity</span> measurements led us to identify 145 RV constant stars with radial <span class="hlt">velocities</span> varying by less than 1 km s-1. Seventy-eight stars show significant RV scatter, while nine stars show a composite spectrum</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JAP...122c5702G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JAP...122c5702G"><span>Ab initio <span class="hlt">velocity-field</span> curves in monoclinic β-Ga2O3</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ghosh, Krishnendu; Singisetti, Uttam</p> <p>2017-07-01</p> <p>We investigate the high-<span class="hlt">field</span> transport in monoclinic β-Ga2O3 using a combination of ab initio calculations and full band Monte Carlo (FBMC) simulation. Scattering rate calculation and the final state selection in the FBMC simulation use complete wave-vector (both electron and phonon) and crystal direction dependent electron phonon interaction (EPI) elements. We propose and implement a semi-coarse version of the Wannier-Fourier interpolation method [Giustino et al., Phys. Rev. B 76, 165108 (2007)] for short-range non-polar optical phonon (EPI) elements in order to ease the computational requirement in FBMC simulation. During the interpolation of the EPI, the inverse Fourier sum over the real-space electronic grids is done on a coarse mesh while the unitary rotations are done on a fine mesh. This paper reports the high <span class="hlt">field</span> transport in monoclinic β-Ga2O3 with deep insight into the contribution of electron-phonon interactions and <span class="hlt">velocity-field</span> characteristics for electric <span class="hlt">fields</span> ranging up to 450 kV/cm in different crystal directions. A peak <span class="hlt">velocity</span> of 2 × 107 cm/s is estimated at an electric <span class="hlt">field</span> of 200 kV/cm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1994JGR....9911417G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1994JGR....9911417G"><span><span class="hlt">Field</span>-aligned Poynting flux <span class="hlt">observations</span> in the high-latitude ionosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gary, J. B.; Heelis, R. A.; Hanson, W. B.; Slavin, J. A.</p> <p>1994-06-01</p> <p>We have used data from Dynamics Explorer 2 to investigate the rate of conversion of electromagnetic energy into both thermal and bulk flow particle kinetic energy in the high-latitude ionosphere. The flux tube integrated conversion rate E.J can be determined from spacecraft measurements of the electric and magnetic <span class="hlt">field</span> vectors by deriving the <span class="hlt">field</span>-aligned Poynting flux, S∥=S.B0, where B0 is in the direction of the geomagnetic <span class="hlt">field</span>. Determination of the Poynting flux from satellite <span class="hlt">observations</span> is critically dependent upon the establishment of accurate values of the <span class="hlt">fields</span> and is especially sensitive to errors in the baseline (unperturbed) geomagnetic <span class="hlt">field</span>. We discuss our treatment of the data in some detail, particularly in regard to systematically correcting the measured magnetic <span class="hlt">field</span> to account for attitude changes and model deficiencies. S∥ can be used to identify the relative strengths of the magnetosphere and thermospheric winds as energy drivers and we present <span class="hlt">observations</span> demonstrating the dominance of each of these. Dominance of the magnetospheric driver is indicated by S∥ directed into the ionosphere. Electromagnetic energy is delivered to and dissipated within the region. Dominance of the neutral wind requires that the conductivity weighted neutral wind speed in the direction of the ion drift be larger than the ion drift, resulting in <span class="hlt">observations</span> of an upward directed Poynting flux. Electromagnetic energy is generated within the ionospheric region in this case. We also present <span class="hlt">observations</span> of a case where the neutral atmosphere motion may be reaching a state of sustained bulk flow <span class="hlt">velocity</span> as evidenced by very small Poynting flux in the presence of large electric <span class="hlt">fields</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910049118&hterms=wall+turbulence&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dwall%2Bturbulence','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910049118&hterms=wall+turbulence&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dwall%2Bturbulence"><span><span class="hlt">Velocity</span> and pressure <span class="hlt">fields</span> associated with near-wall turbulence structures</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Johansson, Arne V.; Alfredsson, P. Henrik; Kim, John</p> <p>1990-01-01</p> <p>Computer generated databases containing <span class="hlt">velocity</span> and pressure <span class="hlt">fields</span> in three-dimensional space at a sequence of time-steps, were used for the investigation of near-wall turbulence structures, their space-time evolution, and their associated pressure <span class="hlt">fields</span>. The main body of the results were obtained from simulation data for turbulent channel flow at a Reynolds number of 180 (based on half-channel height and friction <span class="hlt">velocity</span>) with a grid of 128 x 129 x and 128 points. The flow was followed over a total time of 141 viscous time units. Spanwise centering of the detected structures was found to be essential in order to obtain a correct magnitude of the associated Reynolds stress contribution. A positive wall-pressure peak is found immediately beneath the center of the structure. The maximum amplitude of the pressure pattern was, however, found in the buffer region at the center of the shear-layer. It was also found that these flow structures often reach a maximum strength in connection with an asymmetric spanwise motion, which motivated the construction of a conditional sampling scheme that preserved this asymmetry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110011011','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110011011"><span>Whistler Waves Driven by Anisotropic Strahl <span class="hlt">Velocity</span> Distributions: Cluster <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>Vinas, A.F.; Gurgiolo, C.; Nieves-Chinchilla, T.; Gary, S. P.; Goldstein, M. L.</p> <p>2010-01-01</p> <p><span class="hlt">Observed</span> properties of the strahl using high resolution 3D electron <span class="hlt">velocity</span> 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 <span class="hlt">velocity</span> distributions in the solar wind through pitch-angle scattering</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PASP..129h4201F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PASP..129h4201F"><span>IN-SYNC VI. Identification and Radial <span class="hlt">Velocity</span> Extraction for 100+ Double-Lined Spectroscopic Binaries in the APOGEE/IN-SYNC <span class="hlt">Fields</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fernandez, M. A.; Covey, Kevin R.; De Lee, Nathan; Chojnowski, S. Drew; Nidever, David; Ballantyne, Richard; Cottaar, Michiel; Da Rio, Nicola; Foster, Jonathan B.; Majewski, Steven R.; Meyer, Michael R.; Reyna, A. M.; Roberts, G. W.; Skinner, Jacob; Stassun, Keivan; Tan, Jonathan C.; Troup, Nicholas; Zasowski, Gail</p> <p>2017-08-01</p> <p>We present radial <span class="hlt">velocity</span> measurements for 70 high confidence, and 34 potential binary systems in <span class="hlt">fields</span> containing the Perseus Molecular Cloud, Pleiades, NGC 2264, and the Orion A star-forming region. Eighteen of these systems have been previously identified as binaries in the literature. Candidate double-lined spectroscopic binaries (SB2s) are identified by analyzing the cross-correlation functions (CCFs) computed during the reduction of each APOGEE spectrum. We identify sources whose CCFs are well fit as the sum of two Lorentzians as likely binaries, and provide an initial characterization of the system based on the radial <span class="hlt">velocities</span> indicated by that dual fit. For systems <span class="hlt">observed</span> over several epochs, we present mass ratios and systemic <span class="hlt">velocities</span>; for two systems with <span class="hlt">observations</span> on eight or more epochs, and which meet our criteria for robust orbital coverage, we derive initial orbital parameters. The distribution of mass ratios for multi-epoch sources in our sample peaks at q = 1, but with a significant tail toward lower q values. Tables reporting radial <span class="hlt">velocities</span>, systemic <span class="hlt">velocities</span>, and mass ratios are provided online. We discuss future improvements to the radial <span class="hlt">velocity</span> extraction method we employ, as well as limitations imposed by the number of epochs currently available in the APOGEE database. The Appendix contains brief notes from the literature on each system in the sample, and more extensive notes for select sources of interest.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMEP21A0884M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMEP21A0884M"><span>Three-dimensional simulation of the motion of a single particle under a simulated turbulent <span class="hlt">velocity</span> <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Moreno-Casas, P. A.; Bombardelli, F. A.</p> <p>2015-12-01</p> <p>A 3D Lagrangian particle tracking model is coupled to a 3D channel <span class="hlt">velocity</span> <span class="hlt">field</span> to simulate the saltation motion of a single sediment particle moving in saltation mode. The turbulent <span class="hlt">field</span> is a high-resolution three dimensional <span class="hlt">velocity</span> <span class="hlt">field</span> that reproduces a by-pass transition to turbulence on a flat plate due to free-stream turbulence passing above de plate. In order to reduce computational costs, a decoupled approached is used, i.e., the turbulent flow is simulated independently from the tracking model, and then used to feed the 3D Lagrangian particle model. The simulations are carried using the point-particle approach. The particle tracking model contains three sub-models, namely, particle free-flight, a post-collision <span class="hlt">velocity</span> and bed representation sub-models. The free-flight sub-model considers the action of the following forces: submerged weight, non-linear drag, lift, virtual mass, Magnus and Basset forces. The model also includes the effect of particle angular <span class="hlt">velocity</span>. The post-collision <span class="hlt">velocities</span> are obtained by applying conservation of angular and linear momentum. The complete model was validated with experimental results from literature within the sand range. Results for particle <span class="hlt">velocity</span> time series and distribution of particle turbulent intensities are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ResPh...8.1104H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ResPh...8.1104H"><span>Flow of nanofluid by nonlinear stretching <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>Hayat, Tasawar; Rashid, Madiha; Alsaedi, Ahmed; Ahmad, Bashir</p> <p>2018-03-01</p> <p>Main objective in this article is to model and analyze the nanofluid flow induced by curved surface with nonlinear stretching <span class="hlt">velocity</span>. Nanofluid comprises water and silver. Governing problem is solved by using homotopy analysis method (HAM). Induced magnetic <span class="hlt">field</span> for low magnetic Reynolds number is not entertained. Development of convergent series solutions for <span class="hlt">velocity</span> and skin friction coefficient is successfully made. Pressure in the boundary layer flow by curved stretching surface cannot be ignored. It is found that magnitude of power-law index parameter increases for pressure distibutions. Magnitude of radius of curvature reduces for pressure <span class="hlt">field</span> while opposite trend can be <span class="hlt">observed</span> for <span class="hlt">velocity</span>.</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 <span class="hlt">field</span>-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('http://adsabs.harvard.edu/abs/2008AGUFM.G31A0638C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.G31A0638C"><span>Predicting present-day rates of glacial isostatic adjustment using a smoothed GPS <span class="hlt">velocity</span> <span class="hlt">field</span> for the reconciliation of NAD83 reference frames in Canada</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Craymer, M. R.; Henton, J. A.; Piraszewski, M.</p> <p>2008-12-01</p> <p>Glacial isostatic adjustment following the last glacial period is the dominant source of crustal deformation in Canada east of the Rocky Mountains. The present-day vertical component of motion associated with this process may exceed 1 cm/y and is being directly measured with the Global Positioning System (GPS). A consequence of this steady deformation is that high accuracy coordinates at one epoch may not be compatible with those at another epoch. For example, modern precise point positioning (PPP) methods provide coordinates at the epoch of <span class="hlt">observation</span> while NAD83, the officially adopted reference frame in Canada and the U.S., is expressed at some past reference epoch. The PPP positions are therefore incompatible with coordinates in such a realization of the reference frame and need to be propagated back to the frame's reference epoch. Moreover, the realizations of NAD83 adopted by the provincial geodetic agencies in Canada are referenced to different coordinate epochs; either 1997.0 or 2002.0. Proper comparison of coordinates between provinces therefore requires propagating them from one reference epoch to another. In an effort to reconcile PPP results and different realizations of NAD83, we empirically represent crustal deformation throughout Canada using a <span class="hlt">velocity</span> <span class="hlt">field</span> based solely on high accuracy continuous and episodic GPS <span class="hlt">observations</span>. The continuous <span class="hlt">observations</span> from 2001 to 2007 were obtained from nearly 100 permanent GPS stations, predominately operated by Natural Resources Canada (NRCan) and provincial geodetic agencies. Many of these sites are part of the International GNSS Service (IGS) global network. Episodic <span class="hlt">observations</span> from 1994 to 2006 were obtained from repeated occupations of the Canadian Base Network (CBN), which consists of approximately 160 stable pillar-type monuments across the entire country. The CBN enables a much denser spatial sampling of crustal motions although coverage in the far north is still rather sparse. NRCan solutions of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..1715577T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..1715577T"><span>A simple measuring technique of surface flow <span class="hlt">velocity</span> to analyze the behavior of <span class="hlt">velocity</span> <span class="hlt">fields</span> in hydraulic engineering applications.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tellez, Jackson; Gomez, Manuel; Russo, Beniamino; Redondo, Jose M.</p> <p>2015-04-01</p> <p>An important achievement in hydraulic engineering is the proposal and development of new techniques for the measurement of <span class="hlt">field</span> <span class="hlt">velocities</span> in hydraulic problems. The technological advances in digital cameras with high resolution and high speed found in the market, and the advances in digital image processing techniques now provides a tremendous potential to measure and study the behavior of the water surface flows. This technique was applied at the Laboratory of Hydraulics at the Technical University of Catalonia - Barcelona Tech to study the 2D <span class="hlt">velocity</span> <span class="hlt">fields</span> in the vicinity of a grate inlet. We used a platform to test grate inlets capacity with dimensions of 5.5 m long and 4 m wide allowing a zone of useful study of 5.5m x 3m, where the width is similar of the urban road lane. The platform allows you to modify the longitudinal slopes from 0% to 10% and transversal slope from 0% to 4%. Flow rates can arrive to 200 l/s. In addition a high resolution camera with 1280 x 1024 pixels resolution with maximum speed of 488 frames per second was used. A novel technique using particle image velocimetry to measure surface flow <span class="hlt">velocities</span> has been developed and validated with the experimental data from the grate inlets capacity. In this case, the proposed methodology can become a useful tools to understand the <span class="hlt">velocity</span> <span class="hlt">fields</span> of the flow approaching the inlet where the traditional measuring equipment have serious problems and limitations. References DigiFlow User Guide. (2012), (June). Russo, B., Gómez, M., & Tellez, J. (2013). Methodology to Estimate the Hydraulic Efficiency of Nontested Continuous Transverse Grates. Journal of Irrigation and Drainage Engineering, 139(10), 864-871. doi:10.1061/(ASCE)IR.1943-4774.0000625 Teresa Vila (1), Jackson Tellez (1), Jesus Maria Sanchez (2), Laura Sotillos (1), Margarita Diez (3, 1), and J., & (1), M. R. (2014). Diffusion in fractal wakes and convective thermoelectric flows. Geophysical Research Abstracts - EGU General Assembly 2014</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70185053','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70185053"><span>Does resolution of flow <span class="hlt">field</span> <span class="hlt">observation</span> influence apparent habitat use and energy expenditure in juvenile coho salmon?</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Tullos, Desiree D.; Walter, Cara; Dunham, Jason B.</p> <p>2016-01-01</p> <p>This study investigated how the resolution of <span class="hlt">observation</span> influences interpretation of how fish, juvenile Coho Salmon (Oncorhynchus kisutch), exploit the hydraulic environment in streams. Our objectives were to evaluate how spatial resolution of the flow <span class="hlt">field</span> <span class="hlt">observation</span> influenced: (1) the <span class="hlt">velocities</span> considered to be representative of habitat units; (2) patterns of use of the hydraulic environment by fish; and (3) estimates of energy expenditure. We addressed these objectives using <span class="hlt">observations</span> within a 1:1 scale physical model of a full-channel log jam in an outdoor experimental stream. <span class="hlt">Velocities</span> were measured with Acoustic Doppler Velocimetry at a 10 cm grid spacing, whereas fish locations and tailbeat frequencies were documented over time using underwater videogrammetry. Results highlighted that resolution of <span class="hlt">observation</span> did impact perceived habitat use and energy expenditure, as did the location of measurement within habitat units and the use of averaging to summarize <span class="hlt">velocities</span> within a habitat unit. In this experiment, the range of <span class="hlt">velocities</span> and energy expenditure estimates increased with coarsening resolution (grid spacing from 10 to 100 cm), reducing the likelihood of measuring the <span class="hlt">velocities</span> locally experienced by fish. In addition, the coarser resolutions contributed to fish appearing to select <span class="hlt">velocities</span> that were higher than what was measured at finer resolutions. These findings indicate the need for careful attention to and communication of resolution of <span class="hlt">observation</span> in investigating the hydraulic environment and in determining the habitat needs and bioenergetics of aquatic biota.</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> <span class="hlt">field</span> 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> gradients 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> gradients 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> <span class="hlt">field</span> but not the instantaneous <span class="hlt">velocity</span> <span class="hlt">field</span> (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> <span class="hlt">field</span> and the average rate of moment release in the lithosphere. Examples with synthetic fault networks demonstrate that <span class="hlt">velocity</span> gradients 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('http://adsabs.harvard.edu/abs/2015AGUFM.C21B0741W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.C21B0741W"><span>How can we Optimize Global Satellite <span class="hlt">Observations</span> of Glacier <span class="hlt">Velocity</span> and Elevation Changes?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Willis, M. J.; Pritchard, M. E.; Zheng, W.</p> <p>2015-12-01</p> <p>We have started a global compilation of glacier surface elevation change rates measured by altimeters and differencing of Digital Elevation Models and glacier <span class="hlt">velocities</span> 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 <span class="hlt">velocities</span> 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 <span class="hlt">velocities</span> 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 <span class="hlt">velocities</span>, 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 <span class="hlt">observations</span> must be made in order to provide quality satellite-derived <span class="hlt">velocity</span> and ice elevation <span class="hlt">observations</span> 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 <span class="hlt">observations</span> from the global constellation of satellite missions to record important changes to glacier elevations and <span class="hlt">velocities</span> worldwide.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998PhRvB..58.7230K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998PhRvB..58.7230K"><span>Distribution of electromagnetic <span class="hlt">field</span> and group <span class="hlt">velocities</span> in two-dimensional periodic systems with dissipative metallic components</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kuzmiak, Vladimir; Maradudin, Alexei A.</p> <p>1998-09-01</p> <p>We study the distribution of the electromagnetic <span class="hlt">field</span> of the eigenmodes and corresponding group <span class="hlt">velocities</span> associated with the photonic band structures of two-dimensional periodic systems consisting of an array of infinitely long parallel metallic rods whose intersections with a perpendicular plane form a simple square lattice. We consider both nondissipative and lossy metallic components characterized by a complex frequency-dependent dielectric function. Our analysis is based on the calculation of the complex photonic band structure obtained by using a modified plane-wave method that transforms the problem of solving Maxwell's equations into the problem of diagonalizing an equivalent non-Hermitian matrix. In order to investigate the nature and the symmetry properties of the eigenvectors, which significantly affect the optical properties of the photonic lattices, we evaluate the associated <span class="hlt">field</span> distribution at the high symmetry points and along high symmetry directions in the two-dimensional first Brillouin zone of the periodic system. By considering both lossless and lossy metallic rods we study the effect of damping on the spatial distribution of the eigenvectors. Then we use the Hellmann-Feynman theorem and the eigenvectors and eigenfrequencies obtained from a photonic band-structure calculation based on a standard plane-wave approach applied to the nondissipative system to calculate the components of the group <span class="hlt">velocities</span> associated with individual bands as functions of the wave vector in the first Brillouin zone. From the group <span class="hlt">velocity</span> of each eigenmode the flow of energy is examined. The results obtained indicate a strong directional dependence of the group <span class="hlt">velocity</span>, and confirm the experimental <span class="hlt">observation</span> that a photonic crystal is a potentially efficient tool in controlling photon propagation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19950015537','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19950015537"><span>The <span class="hlt">velocity</span> and vorticity <span class="hlt">fields</span> of the turbulent near wake of a circular cylinder</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wallace, James; Ong, Lawrence; Moin, Parviz</p> <p>1995-01-01</p> <p>The purpose of this research is to provide a detailed experimental database of <span class="hlt">velocity</span> and vorticity statistics in the very near wake (x/d less than 10) of a circular cylinder at Reynolds number of 3900. This study has determined that estimations of the streamwise <span class="hlt">velocity</span> component in flow <span class="hlt">fields</span> with large nonzero cross-stream components are not accurate. Similarly, X-wire measurements of the u and v <span class="hlt">velocity</span> components in flows containing large w are also subject to the errors due to binormal cooling. Using the look-up table (LUT) technique, and by calibrating the X-wire probe used here to include the range of expected angles of attack (+/- 40 deg), accurate X-wire measurements of instantaneous u and v <span class="hlt">velocity</span> components in the very near wake region of a circular cylinder has been accomplished. The approximate two-dimensionality of the present flow <span class="hlt">field</span> was verified with four-wire probe measurements, and to some extent the spanwise correlation measurements with the multisensor rake. Hence, binormal cooling errors in the present X-wire measurements are small.</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/70018410','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70018410"><span>Crustal <span class="hlt">velocity</span> <span class="hlt">field</span> near the big bend of California's San Andreas fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Snay, R.A.; Cline, M.W.; Philipp, C.R.; Jackson, D.D.; Feng, Y.; Shen, Z.-K.; Lisowski, M.</p> <p>1996-01-01</p> <p>We use geodetic data spanning the 1920-1992 interval to estimate the horizontal <span class="hlt">velocity</span> <span class="hlt">field</span> near the big bend segment of California's San Andreas fault (SAF). More specifically, we estimate a horizontal <span class="hlt">velocity</span> vector for each node of a two-dimensional grid that has a 15-min-by-15-min mesh and that extends between latitudes 34.0??N and 36.0??N and longitudes 117.5??W and 120.5??W. For this estimation process, we apply bilinear interpolation to transfer crustal deformation information from geodetic sites to the grid nodes. The data include over a half century of triangulation measurements, over two decades of repeated electronic distance measurements, a decade of repeated very long baseline interferometry measurements, and several years of Global Positioning System measurements. Magnitudes for our estimated <span class="hlt">velocity</span> vectors have formal standard errors ranging from 0.7 to 6.8 mm/yr. Our derived <span class="hlt">velocity</span> <span class="hlt">field</span> shows that (1) relative motion associated with the SAF exceeds 30 mm/yr and is distributed on the Earth's surface across a band (> 100 km wide) that is roughly centered on this fault; (2) when <span class="hlt">velocities</span> are expressed relative to a fixed North America plate, the motion within our primary study region has a mean orientation of N44??W ?? 2?? and the surface trace of the SAF is congruent in shape to nearby contours of constant speed yet this trace is oriented between 5?? and 10?? counterclockwise relative to these contours; and (3) large strain rates (shear rates > 150 nrad/yr and/or areal dilatation rates < -150 nstr/yr) exist near the Garlock fault, near the White Wolf fault, and in the Ventura basin.</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://hdl.handle.net/2060/19810004457','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19810004457"><span><span class="hlt">Observations</span> of the magnetic <span class="hlt">field</span> and plasma flow in Jupiter's magnetosheath</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lepping, R. P.; Burlaga, L. F.; Klein, L. W.; Jessen, J. M.; Goodrich, G. C.</p> <p>1980-01-01</p> <p>Large scale (many minutes to 10 hours) magnetic <span class="hlt">field</span> structures consisting predominantly of nearly north-south <span class="hlt">field</span> direction were discovered in Jupiter's magnetosheath from the data of Voyagers 1 and 2 and Pioneer 10 during their outbound encounter trajectories. The Voyager 2 data, and that of Voyager 1 to a lesser extent, show evidence of a quasi-period of 10 hours (and occasionally 5 hours) for these structures. The north-south components of the <span class="hlt">field</span> and plasma <span class="hlt">velocity</span> were strongly correlated in the outbound magnetosheath as <span class="hlt">observed</span> by Voyagers 1 and 2, and the components orthogonal to the north-south direction showed weak correlations. For both Voyager encounters the sense (positive and negative) of the north-south correlations were directly related to the direction of the ecliptic plane component of the interplanetary magnetic <span class="hlt">field</span> using the <span class="hlt">field</span> and plasma measurements of the non-encountering spacecraft.</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 <span class="hlt">fields</span> is to focus on the role of spokes in the process of electronmore » energization.« less</p> </li> <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 <span class="hlt">field</span> <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/2016usc..confE..93S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016usc..confE..93S"><span>HMI Measured Doppler <span class="hlt">Velocity</span> Contamination from the SDO Orbit <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>Scherrer, Phil; HMI Team</p> <p>2016-10-01</p> <p>The Problem: The SDO satellite is in an inclined Geo-sync orbit which allows uninterrupted views of the Sun nearly 98% of the time. This orbit has a <span class="hlt">velocity</span> of about 3,500 m/s with the solar line-of-sight component varying with time of day and time of year. Due to remaining calibration errors in wavelength filters the orbit <span class="hlt">velocity</span> leaks into the line-of-sight solar <span class="hlt">velocity</span> and magnetic <span class="hlt">field</span> measurements. Since the same model of the filter is used in the Milne-Eddington inversions used to generate the vector magnetic <span class="hlt">field</span> data, the orbit <span class="hlt">velocity</span> also contaminates the vector magnetic products. These errors contribute 12h and 24h variations in most HMI data products and are known as the 24-hour problem. Early in the mission we made a patch to the calibration that corrected the disk mean <span class="hlt">velocity</span>. The resulting LOS <span class="hlt">velocity</span> has been used for helioseismology with no apparent problems. The <span class="hlt">velocity</span> signal has about a 1% scale error that varies with time of day and with <span class="hlt">velocity</span>, i.e. it is non-linear for large <span class="hlt">velocities</span>. This causes leaks into the LOS <span class="hlt">field</span> (which is simply the difference between <span class="hlt">velocity</span> measured in LCP and RCP rescaled for the Zeeman splitting). This poster reviews the measurement process, shows examples of the problem, and describes recent work at resolving the issues. Since the errors are in the filter characterization it makes most sense to work first on the LOS data products since they, unlike the vector products, are directly and simply related to the filter profile without assumptions on the solar atmosphere, filling factors, etc. Therefore this poster is strictly limited to understanding how to better understand the filter profiles as they vary across the <span class="hlt">field</span> and with time of day and time in years resulting in <span class="hlt">velocity</span> errors of up to a percent and LOS <span class="hlt">field</span> estimates with errors up to a few percent (of the standard LOS magnetograph method based on measuring the differences in wavelength of the line centroids in LCP and RCP light). We</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20030003706','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20030003706"><span>Measurement of Correlation Between Flow Density, <span class="hlt">Velocity</span>, and Density*<span class="hlt">velocity</span>(sup 2) with Far <span class="hlt">Field</span> Noise in High Speed Jets</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Panda, Jayanta; Seasholtz, Richard G.; Elam, Kristie A.</p> <p>2002-01-01</p> <p>To locate noise sources in high-speed jets, the sound pressure fluctuations p', measured at far <span class="hlt">field</span> locations, were correlated with each of radial <span class="hlt">velocity</span> v, density rho, and phov(exp 2) fluctuations measured from various points in jet plumes. The experiments follow the cause-and-effect method of sound source identification, where <rhov(exp 2)-p'> correlation is related to the first, and <rho-p'> correlation to the second source terms of Lighthill's equation. Three fully expanded, unheated plumes of Mach number 0.95, 1.4 and 1.8 were studied for this purpose. The <span class="hlt">velocity</span> and density fluctuations were measured simultaneously using a recently developed, non-intrusive, point measurement technique based on molecular Rayleigh scattering. It was <span class="hlt">observed</span> that along the jet centerline the density fluctuation spectra S(sub rho) have different shapes than the radial <span class="hlt">velocity</span> spectra S(sub v), while data obtained from the peripheral shear layer show similarity between the two spectra. Density fluctuations in the jet showed significantly higher correlation, than either rhov(sub 2) or v fluctuations. It is found that a single point <rho-p'> correlation from the peak sound emitting region at the end of the potential core can account for nearly 10% of all noise at 30 to the jet axis. The <phov(exp 2)-p'> correlation, representing the effectiveness of a longitudinal quadrupole in generating noise 90 to the jet axis, is found to be zero within experimental uncertainty. In contrast rhov(exp 2) fluctuations were better correlated with sound pressure fluctuation at the 30 location. The strongest source of sound is found to lie at the centerline and beyond the end of potential core.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeoRL..45.1271Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeoRL..45.1271Y"><span>Excitation of O+ Band EMIC Waves Through H+ Ring <span class="hlt">Velocity</span> Distributions: Van Allen Probe <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>Yu, Xiongdong; Yuan, Zhigang; Huang, Shiyong; Yao, Fei; Wang, Dedong; Funsten, Herbert O.; Wygant, John R.</p> <p>2018-02-01</p> <p>A typical case of electromagnetic ion cyclotron (EMIC) emissions with both He+ band and O+ band waves was <span class="hlt">observed</span> by Van Allen Probe A on 14 July 2014. These emissions occurred in the morning sector on the equator inside the plasmasphere, in which region O+ band EMIC waves prefer to appear. Through property analysis of these emissions, it is found that the He+ band EMIC waves are linearly polarized and propagating quasi-parallelly along the background magnetic <span class="hlt">field</span>, while the O+ band ones are of linear and left-hand polarization and propagating obliquely with respect to the background magnetic <span class="hlt">field</span>. Using the in situ <span class="hlt">observations</span> of plasma environment and particle data, excitation of these O+ band EMIC waves has been investigated with the linear growth theory. The calculated linear growth rate shows that these O+ band EMIC waves can be locally excited by ring current protons with ring <span class="hlt">velocity</span> distributions. The comparison of the <span class="hlt">observed</span> wave spectral intensity and the calculated growth rate suggests that the density of H+ rings providing the free energy for the instability has decreased after the wave grows. Therefore, this paper provides a direct <span class="hlt">observational</span> evidence to the excitation mechanism of O+ band EMIC waves: ring current protons with ring distributions provide the free energy supporting the instability in the presence of rich O+ in the plasmasphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22649747-field-dependence-electron-drift-velocity-along-hexagonal-axis-sic','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22649747-field-dependence-electron-drift-velocity-along-hexagonal-axis-sic"><span><span class="hlt">Field</span> dependence of the electron drift <span class="hlt">velocity</span> along the hexagonal axis of 4H-SiC</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, P. A., E-mail: Pavel.Ivanov@mail.ioffe.ru; Potapov, A. S.; Samsonova, T. P.</p> <p></p> <p>The forward current–voltage characteristics of mesa-epitaxial 4H-SiC Schottky diodes are measured in high electric <span class="hlt">fields</span> (up to 4 × 10{sup 5} V/cm) in the n-type base region. A semi-empirical formula for the <span class="hlt">field</span> dependence of the electron drift <span class="hlt">velocity</span> in 4H-SiC along the hexagonal axis of the crystal is derived. It is shown that the saturated drift <span class="hlt">velocity</span> is (1.55 ± 0.05) × 10{sup 7} cm/s in electric <span class="hlt">fields</span> higher than 2 × 10{sup 5} V/cm.</p> </li> <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('https://www.ncbi.nlm.nih.gov/pubmed/18547928','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18547928"><span>Assessment of <span class="hlt">velocity</span> <span class="hlt">fields</span> through open-channel flows with an empiric law.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bardiaux, J B; Vazquez, J; Mosé, R</p> <p>2008-01-01</p> <p>Most sewer managers are currently confronted with the evaluation of the water discharges, that flow through their networks or go to the discharge system, i.e. rivers in the majority of cases. In this context, the Urban Hydraulic Systems laboratory of the ENGEES is working on the relation between <span class="hlt">velocity</span> <span class="hlt">fields</span> and metrology assessment through a partnership with the Fluid and Solid Mechanics Institute of Strasbourg (IMFS). The responsibility is clearly to transform a <span class="hlt">velocity</span> profile measurement, given by a Doppler sensor developed by the IMFS team, into a water discharge evaluation. The <span class="hlt">velocity</span> distribution in a cross section of the flow in a channel has attracted the interests of many researchers over the years, due to its practical applications. In the case of free surface flows in narrow open channels the maximum <span class="hlt">velocity</span> is below the free surface. This phenomenon, usually called "dip-phenomenon", amongst other things, raises the problem of the area explored in the section of measurements. The work presented here tries to create a simple relation making possible to associate the flow with the <span class="hlt">velocity</span> distribution. This step allows to insert the sensor position into the flow calculation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3345833','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3345833"><span>GPS <span class="hlt">Velocity</span> and Strain Rate <span class="hlt">Fields</span> in Southwest Anatolia from Repeated GPS Measurements</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Erdoğan, Saffet; Şahin, Muhammed; Tiryakioğlu, İbrahim; Gülal, Engin; Telli, Ali Kazım</p> <p>2009-01-01</p> <p>Southwestern Turkey is a tectonically active area. To determine kinematics and strain distribution in this region, a GPS network of sixteen stations was established. We have used GPS <span class="hlt">velocity</span> <span class="hlt">field</span> data for southwest Anatolia from continuous measurements covering the period 2003 to 2006 to estimate current crustal deformation of this tectonically active region. GPS data were processed using GAMIT/GLOBK software and <span class="hlt">velocity</span> and strain rate <span class="hlt">fields</span> were estimated in the study area. The measurements showed <span class="hlt">velocities</span> of 15–30 mm/yr toward the southwest and strain values up to 0.28–8.23×10−8. Results showed that extension has been determined in the Burdur-Isparta region. In this study, all of strain data reveal an extensional neotectonic regime through the northeast edge of the Isparta Angle despite the previously reported compressional neotectonic regime. Meanwhile, results showed some small differences relatively with the 2006 model of Reilinger et al. As a result, active tectonic movements, in agreement with earthquake fault plane solutions showed important activity. PMID:22573998</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28808341','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28808341"><span>Tuning the Fermi <span class="hlt">velocity</span> in Dirac materials with an electric <span class="hlt">field</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Díaz-Fernández, A; Chico, Leonor; González, J W; Domínguez-Adame, F</p> <p>2017-08-14</p> <p>Dirac materials are characterized by energy-momentum relations that resemble those of relativistic massless particles. Commonly denominated Dirac cones, these dispersion relations are considered to be their essential feature. These materials comprise quite diverse examples, such as graphene and topological insulators. Band-engineering techniques should aim to a full control of the parameter that characterizes the Dirac cones: the Fermi <span class="hlt">velocity</span>. We propose a general mechanism that enables the fine-tuning of the Fermi <span class="hlt">velocity</span> in Dirac materials in a readily accessible way for experiments. By embedding the sample in a uniform electric <span class="hlt">field</span>, the Fermi <span class="hlt">velocity</span> is substantially modified. We first prove this result analytically, for the surface states of a topological insulator/semiconductor interface, and postulate its universality in other Dirac materials. Then we check its correctness in carbon-based Dirac materials, namely graphene nanoribbons and nanotubes, thus showing the validity of our hypothesis in different Dirac systems by means of continuum, tight-binding and ab-initio calculations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20070031919','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20070031919"><span><span class="hlt">Observations</span> of the Ion Signatures of Double Merging and the Formation of Newly Closed <span class="hlt">Field</span> Lines</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Chandler, Michael O.; Avanov, Levon A.; Craven, Paul D.</p> <p>2007-01-01</p> <p><span class="hlt">Observations</span> from the Polar spacecraft, taken during a period of northward interplanetary magnetic <span class="hlt">field</span> (IMF) show magnetosheath ions within the magnetosphere with <span class="hlt">velocity</span> distributions resulting from multiple merging sites along the same <span class="hlt">field</span> line. The <span class="hlt">observations</span> from the TIDE instrument show two separate ion energy-time dispersions that are attributed to two widely separated (-20Re) merging sites. Estimates of the initial merging times show that they occurred nearly simultaneously (within 5 minutes.) Along with these populations, cold, ionospheric ions were <span class="hlt">observed</span> counterstreaming along the <span class="hlt">field</span> lines. The presence of such ions is evidence that these <span class="hlt">field</span> lines are connected to the ionosphere on both ends. These results are consistent with the hypothesis that double merging can produce closed <span class="hlt">field</span> lines populated by solar wind plasma. While the merging sites cannot be unambiguously located, the <span class="hlt">observations</span> and analyses favor one site poleward of the northern cusp and a second site at low latitudes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22679664-understanding-redshift-space-distortions-density-weighted-peculiar-velocity','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22679664-understanding-redshift-space-distortions-density-weighted-peculiar-velocity"><span>Understanding redshift space distortions in density-weighted peculiar <span class="hlt">velocity</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>Sugiyama, Naonori S.; Okumura, Teppei; Spergel, David N., E-mail: nao.s.sugiyama@gmail.com, E-mail: teppei.oku@gmail.com, E-mail: dns@astro.princeton.edu</p> <p>2016-07-01</p> <p><span class="hlt">Observations</span> of the kinetic Sunyaev-Zel'dovich (kSZ) effect measure the density-weighted <span class="hlt">velocity</span> <span class="hlt">field</span>, a potentially powerful cosmological probe. This paper presents an analytical method to predict the power spectrum and two-point correlation function of the density-weighted <span class="hlt">velocity</span> in redshift space, the direct <span class="hlt">observables</span> in kSZ surveys. We show a simple relation between the density power spectrum and the density-weighted <span class="hlt">velocity</span> power spectrum that holds for both dark matter and halos. Using this relation, we can then extend familiar perturbation expansion techniques to the kSZ power spectrum. One of the most important features of density-weighted <span class="hlt">velocity</span> statistics in redshift space is themore » change in sign of the cross-correlation between the density and density-weighted <span class="hlt">velocity</span> at mildly small scales due to nonlinear redshift space distortions. Our model can explain this characteristic feature without any free parameters. As a result, our results can precisely predict the non-linear behavior of the density-weighted <span class="hlt">velocity</span> <span class="hlt">field</span> in redshift space up to ∼ 30 h {sup -1} Mpc for dark matter particles at the redshifts of z =0.0, 0.5, and 1.0.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApJ...838...21Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApJ...838...21Z"><span>Evolution of Mass and <span class="hlt">Velocity</span> <span class="hlt">Field</span> in the Cosmic Web: Comparison between Baryonic and Dark Matter</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhu, Weishan; Feng, Long-Long</p> <p>2017-03-01</p> <p>We investigate the evolution of the cosmic web since z = 5 in grid-based cosmological hydrodynamical simulations, focusing on the mass and <span class="hlt">velocity</span> <span class="hlt">fields</span> of both baryonic and cold dark matter. The tidal tensor of density is used as the main method for web identification, with λ th = 0.2-1.2. The evolution trends in baryonic and dark matter are similar, although moderate differences are <span class="hlt">observed</span>. Sheets appear early, and their large-scale pattern may have been set up by z = 3. In terms of mass, filaments supersede sheets as the primary collapsing structures from z ˜ 2-3. Tenuous filaments assembled with each other to form prominent ones at z < 2. In accordance with the construction of the frame of the sheets, the cosmic divergence <span class="hlt">velocity</span>, v div, was already well-developed above 2-3 Mpc by z = 3. Afterwards, the curl <span class="hlt">velocity</span>, v curl, grew dramatically along with the rising of filaments, becoming comparable to v div, for <2-3 Mpc at z = 0. The scaling of v curl can be described by the hierarchical turbulence model. The alignment between the vorticity and the eigenvectors of the shear tensor in the baryonic matter <span class="hlt">field</span> resembles that in the dark matter <span class="hlt">field</span>, and is even moderately stronger between {\\boldsymbol{ω }} and {{\\boldsymbol{e}}}1, and ω and {{\\boldsymbol{e}}}3. Compared with dark matter, there is slightly less baryonic matter found residing in filaments and clusters, and its vorticity developed more significantly below 2-3 Mpc. These differences may be underestimated because of the limited resolution and lack of star formation in our simulation. The impact of the change of dominant structures in overdense regions at z ˜ 2-3 on galaxy formation and evolution is shortly discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.G51C..03D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.G51C..03D"><span>A new GPS <span class="hlt">velocity</span> <span class="hlt">field</span> in the south-western Balkans: insights for continental dynamics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>D'Agostino, N.; Avallone, A.; Duni, L.; Ganas, A.; Georgiev, I.; Jouanne, F.; Koci, R.; Kuka, N.; Metois, M.</p> <p>2017-12-01</p> <p>The Balkans peninsula is an area of active distributed deformation located at the southern boundary of the Eurasian plate. Relatively low strain rates and logistical reasons have so far limited the characterization and definition of the active tectonics and crustal kinematics. The increasing number of GNSS stations belonging to national networks deployed for scientific and cadastral purposes, now provides the opportunity to improve the knowledge of the crustal kinematics in this area and to define a cross-national <span class="hlt">velocity</span> <span class="hlt">field</span> that illuminates the active tectonic deformation. In this work we homogeneously processed the data from the south western Balkans and neighbouring regions using available rinex files from scientific and cadastral networks (ALBPOS, EUREF, HemusNET, ITALPOS, KOPOS, MAKPOS, METRICA, NETGEO, RING, TGREF). In order to analyze and interpret station <span class="hlt">velocities</span> relative to the Eurasia plate and to reduce the common mode signal, we updated the Eurasian terrestrial reference frame described in Métois et al. 2015. Starting from this dataset we present a new GPS <span class="hlt">velocity</span> <span class="hlt">field</span> covering the south western part of the Balkan Peninsula. Using this new <span class="hlt">velocity</span> <span class="hlt">field</span>, we derive the strain rate tensor to analyze the regional style of the deformation. Our results (1) improve the picture of the general southward flow of the crust characterizing the south western Balkans behind the contractional belt at the boundary with Adriatic and (2) provide new key elements for the understanding of continental dynamics in this part of the Eurasian plate boundary.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016WRR....52.1108W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016WRR....52.1108W"><span><span class="hlt">Field</span> assessment of noncontact stream gauging using portable surface <span class="hlt">velocity</span> radars (SVR)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Welber, Matilde; Le Coz, Jérôme; Laronne, Jonathan B.; Zolezzi, Guido; Zamler, Daniel; Dramais, Guillaume; Hauet, Alexandre; Salvaro, Martino</p> <p>2016-02-01</p> <p>The applicability of a portable, commercially available surface <span class="hlt">velocity</span> radar (SVR) for noncontact stream gauging was evaluated through a series of <span class="hlt">field</span>-scale experiments carried out in a variety of sites and deployment conditions. Comparisons with various concurrent techniques showed acceptable agreement with <span class="hlt">velocity</span> profiles, with larger uncertainties close to the banks. In addition to discharge error sources shared with intrusive <span class="hlt">velocity</span>-area techniques, SVR discharge estimates are affected by flood-induced changes in the bed profile and by the selection of a depth-averaged to surface <span class="hlt">velocity</span> ratio, or <span class="hlt">velocity</span> coefficient (α). Cross-sectional averaged <span class="hlt">velocity</span> coefficients showed smaller fluctuations and closer agreement with theoretical values than those computed on individual verticals, especially in channels with high relative roughness. Our findings confirm that α = 0.85 is a valid default value, with a preferred site-specific calibration to avoid underestimation of discharge in very smooth channels (relative roughness ˜ 0.001) and overestimation in very rough channels (relative roughness > 0.05). Theoretically derived and site-calibrated values of α also give accurate SVR-based discharge estimates (within 10%) for low and intermediate roughness flows (relative roughness 0.001 to 0.05). Moreover, discharge uncertainty does not exceed 10% even for a limited number of SVR positions along the cross section (particularly advantageous to gauge unsteady flood flows and very large floods), thereby extending the range of validity of rating curves.</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('https://ntrs.nasa.gov/search.jsp?R=19920064355&hterms=Fluorescence+imaging&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DFluorescence%2Bimaging','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19920064355&hterms=Fluorescence+imaging&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DFluorescence%2Bimaging"><span>Doppler-shifted fluorescence imaging of <span class="hlt">velocity</span> <span class="hlt">fields</span> in supersonic reacting flows</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Allen, M. G.; Davis, S. J.; Kessler, W. J.; Sonnenfroh, D. M.</p> <p>1992-01-01</p> <p>The application of Doppler-shifted fluorescence imaging of <span class="hlt">velocity</span> <span class="hlt">fields</span> in supersonic reacting flows is analyzed. Focussing on fluorescence of the OH molecule in typical H2-air Scramjet flows, the effects of uncharacterized variations in temperature, pressure, and collisional partner composition across the measurement plane are examined. Detailed measurements of the (1,0) band OH lineshape variations in H2-air combustions are used, along with single-pulse and time-averaged measurements of an excimer-pumped dye laser, to predict the performance of a model velocimeter with typical Scramjet flow properties. The analysis demonstrates the need for modification and control of the laser bandshape in order to permit accurate <span class="hlt">velocity</span> measurements in the presence of multivariant flow properties.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DPPC10160Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DPPC10160Z"><span>Heat Transfer to Anode of Arc as Function of Transverse Magnetic <span class="hlt">Field</span> and Lateral Gas Flow <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>Zama, Yoshiyuki; Shiino, Toru; Ishii, Yoko; Maeda, Yoshifumi; Yamamoto, Shinji; Iwao, Toru</p> <p>2016-10-01</p> <p>Gas tungsten arc welding has useful joining technology because of high-energy and high-current characteristics. It can be flexible from the transverse magnetic <span class="hlt">field</span> and lateral gas flow <span class="hlt">velocity</span>. In this case, the weld defect occurs. In this research, the heat transfer to the anode of the arc as a function of the transverse magnetic <span class="hlt">field</span> and lateral gas flow <span class="hlt">velocity</span> is elucidated. That magnetic flux density and lateral gas <span class="hlt">velocity</span> were varied from 0 to 3 mT and 0 to 50?m?s -1, respectively. The axial plasma gas argon flow rates were 3?slm. A transverse magnetic <span class="hlt">field</span> is applied to the arc using Helmholtz coil. The anode is used by a water-cooled copper plate, and the heat transfer is measured by temperature of cooled water. As a result, the arc is deflected by the Lorentz force and lateral gas convection. Thus, the heat transfer to the anode of the arc decreases with increasing the transverse magnetic <span class="hlt">field</span> and lateral gas flow <span class="hlt">velocity</span>. In addition, the heat transfer to the anode changes with different attachments modes. The lateral gas flow causes a convective heat loss from the arc to the chamber walls.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890036918&hterms=barium&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dbarium','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890036918&hterms=barium&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dbarium"><span>Search for auroral belt E-parallel <span class="hlt">fields</span> with high-<span class="hlt">velocity</span> barium ion injections</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Heppner, J. P.; Ledley, B. G.; Miller, M. L.; Marionni, P. A.; Pongratz, M. B.</p> <p>1989-01-01</p> <p>In April 1984, four high-<span class="hlt">velocity</span> shaped-charge Ba(+) injections were conducted from two sounding rockets at 770-975 km over northern Alaska under conditions of active auroral and magnetic disturbance. Spatial ionization (brightness) profiles of high-<span class="hlt">velocity</span> Ba(+) clouds from photometric scans following each release were found to be consistent with the 28-sec theoretical time constant for Ba photoionization determined by Carlsten (1975). These <span class="hlt">observations</span> therefore revealed no evidence of anomalous fast ionization predicted by the Alfven critical <span class="hlt">velocity</span> hypothesis.</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('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('http://adsabs.harvard.edu/abs/2002AAS...201.2306M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002AAS...201.2306M"><span>The Enigmatic Local Hubble Flow: Probing the Nearby Peculiar <span class="hlt">Velocity</span> <span class="hlt">Field</span> with Consistent Distances to Neighboring Galaxies.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mendez, B.; Davis, M.; Newman, J.; Madore, B. F.; Freedman, W. L.; Moustakas, J.</p> <p>2002-12-01</p> <p>The properties of the <span class="hlt">velocity</span> <span class="hlt">field</span> in the local volume (cz < 550 km s-1) have been difficult to constrain due to a lack of a consistent set of galaxy distances. The sparse <span class="hlt">observations</span> available to date suggest a remarkably quiet flow, with little deviation from a pure Hubble law. However, <span class="hlt">velocity</span> <span class="hlt">field</span> models based on the distribution of galaxies in the 1.2 Jy IRAS redshift survey, predict a quadrupolar flow pattern locally with strong infall at the poles of the local Supergalactic plane. In an attempt to resolve this discrepency, we probe the local <span class="hlt">velocity</span> <span class="hlt">field</span> and begin to establish a consistent set of galactic distances. We have obtained images of nearby galaxies in I, V, and B bands from the W.M. Keck Observatory and in F814W and F555W filters from the Hubble Space Telescope. Where these galaxies are well resolved into stars we can use the Tip of the Red Giant Branch (TRGB) as a distance indicator. Using a maximum likelihood analysis to quantitatively measure the I magnitude of the TRGB we determine precise distances to several nearby galaxies. We supplement that dataset with published distances to local galaxies measured using Cepheids, Surface Brightness Fluctuations, and the TRGB. With these data we find that the amplitude of the local flow is roughly half that expected in linear theory and N-body simulations; thus the enigma of cold local flows persists. This work was supported in part by NASA through a grant from the Space Telescope Science Institute and a Predoctoral Fellowship for Minorities from the Ford Foundation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19830043771&hterms=Lenticular&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DLenticular','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830043771&hterms=Lenticular&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DLenticular"><span><span class="hlt">Velocity</span> <span class="hlt">field</span> and physical conditions in the active lenticular galaxy NGC 3998</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Blackman, C. P.; Wilson, A. S.; Ward, M. J.</p> <p>1983-01-01</p> <p>A rotating and expanding flattened distribution of gas is suggested by measurements of the emission line <span class="hlt">velocity</span> <span class="hlt">field</span> for the line elliptical/lenticular galaxy NGC 3998, using seven long slit spectrograms in five position angles. Expanding material kinetic energy values of 10 to the 53rd to 10 to the 54th ergs, together with the flat spectrum radio source and nucleus X-ray emission, indicate pronounced nuclear activity. Spectrophotometry of the galactic nucleus shows emission line strengths typical of shocks rather than of photoionization, and line ratios indicate a postshock temperature of 60,000 K and a preshock density of 25 particles/cu cm. Both the stars and the ionized gas of the galaxy have central <span class="hlt">velocity</span> dispersions of 260 km/s. In view of the high rotational <span class="hlt">velocity</span> of the stars, NGC 3998 is a lenticular rather than elliptical galaxy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22584021-velocity-overshoot-decay-mechanisms-compound-semiconductor-field-effect-transistors-submicron-characteristic-length','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22584021-velocity-overshoot-decay-mechanisms-compound-semiconductor-field-effect-transistors-submicron-characteristic-length"><span><span class="hlt">Velocity</span> overshoot decay mechanisms in compound semiconductor <span class="hlt">field</span>-effect transistors with a submicron characteristic length</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>Jyegal, Jang, E-mail: jjyegal@inu.ac.kr</p> <p></p> <p><span class="hlt">Velocity</span> overshoot is a critically important nonstationary effect utilized for the enhanced performance of submicron <span class="hlt">field</span>-effect devices fabricated with high-electron-mobility compound semiconductors. However, the physical mechanisms of <span class="hlt">velocity</span> overshoot decay dynamics in the devices are not known in detail. Therefore, a numerical analysis is conducted typically for a submicron GaAs metal-semiconductor <span class="hlt">field</span>-effect transistor in order to elucidate the physical mechanisms. It is found that there exist three different mechanisms, depending on device bias conditions. Specifically, at large drain biases corresponding to the saturation drain current (dc) region, the <span class="hlt">velocity</span> overshoot suddenly begins to drop very sensitively due to the onsetmore » of a rapid decrease of the momentum relaxation time, not the mobility, arising from the effect of <span class="hlt">velocity</span>-randomizing intervalley scattering. It then continues to drop rapidly and decays completely by severe mobility reduction due to intervalley scattering. On the other hand, at small drain biases corresponding to the linear dc region, the <span class="hlt">velocity</span> overshoot suddenly begins to drop very sensitively due to the onset of a rapid increase of thermal energy diffusion by electrons in the channel of the gate. It then continues to drop rapidly for a certain channel distance due to the increasing thermal energy diffusion effect, and later completely decays by a sharply decreasing electric <span class="hlt">field</span>. Moreover, at drain biases close to a dc saturation voltage, the mechanism is a mixture of the above two bias conditions. It is suggested that a large secondary-valley energy separation is essential to increase the performance of submicron devices.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.U13B..09K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.U13B..09K"><span>Accessing the inaccessible: making (successful) <span class="hlt">field</span> <span class="hlt">observations</span> at tidewater glacier termini</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kienholz, C.; Amundson, J. M.; Jackson, R. H.; Motyka, R. J.; Nash, J. D.; Sutherland, D.</p> <p>2017-12-01</p> <p>Glaciers terminating in ocean water (tidewater glaciers) show complex dynamic behavior driven predominantly by processes at the ice-ocean interface (sedimentation, erosion, iceberg calving, submarine melting). A quantitative understanding of these processes is required, for example, to better assess tidewater glaciers' fate in our rapidly warming environment. Lacking <span class="hlt">observations</span> close to glacier termini, due to unpredictable risks from calving, hamper this understanding. In an effort to remedy this lack of knowledge, we initiated a large <span class="hlt">field</span>-based effort at LeConte Glacier, southeast Alaska, in 2016. LeConte Glacier is a regional analog for many tidewater glaciers, but better accessible and <span class="hlt">observable</span> and thus an ideal target for our multi-disciplinary effort. Our ongoing campaigns comprise measurements from novel autonomous vessels (temperature, salinity and current) in the immediate proximity of the glacier terminus and additional surveys (including multibeam bathymetry) from boats and moorings in the proglacial fjord. These measurements are complemented by iceberg and glacier <span class="hlt">velocity</span> measurements from time lapse cameras and a portable radar interferometer situated above LeConte Bay. GPS-based <span class="hlt">velocity</span> <span class="hlt">observations</span> and melt measurements are conducted on the glacier. These measurements provide necessary input for process-based understanding and numerical modeling of the glacier and fjord systems. In the presentation, we discuss promising initial results and lessons learned from the campaign.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20080003790','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20080003790"><span><span class="hlt">Observations</span> of a Newly "Captured" Magnetosheath <span class="hlt">Field</span> Line: Evidence for "Double Reconnection"</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Chandler, Michael O.; Avanov, Levon A.; Craven, Paul D.; Mozer, Forrest S.; Moore, Thomas E.</p> <p>2007-01-01</p> <p>We have begun an investigation of the nature of the low-latitude boundary layer in the mid-altitude cusp region using data from the Polar spacecraft. This region has been routinely sampled for about three months each year for the periods 1999-2001 and 2004-2006. The low-to-mid-energy ion instruments frequently <span class="hlt">observed</span> dense, magnetosheath-like plasma deep (in terms of distance from the magnetopause and in invariant latitude) in the magnetosphere. One such case, taken during a period of northward interplanetary magnetic <span class="hlt">field</span> (IMF), shows magnetosheath ions within the magnetosphere with <span class="hlt">velocity</span> distributions resulting from two separate merging sites along the same <span class="hlt">field</span> lines. Cold ionospheric ions were also <span class="hlt">observed</span> counterstreaming along the <span class="hlt">field</span> lines, evidence that these <span class="hlt">field</span> lines were closed. These results are consistent with the hypothesis that double merging can produce closed <span class="hlt">field</span> .lines populated by solar wind plasma. Through the use of individual cases such as this and statistical studies of a broader database we seek to understand the morphology of the LLBL as it projects from the sub-solar region into the cusp. We will present preliminary results of our ongoing study.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhDT.......166D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhDT.......166D"><span>Multiscale Characterization of the Probability Density Functions of <span class="hlt">Velocity</span> and Temperature Increment <span class="hlt">Fields</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>DeMarco, Adam Ward</p> <p></p> <p>The turbulent motions with the atmospheric boundary layer exist over a wide range of spatial and temporal scales and are very difficult to characterize. Thus, to explore the behavior of such complex flow enviroments, it is customary to examine their properties from a statistical perspective. Utilizing the probability density functions of <span class="hlt">velocity</span> and temperature increments, deltau and deltaT, respectively, this work investigates their multiscale behavior to uncover the unique traits that have yet to be thoroughly studied. Utilizing diverse datasets, including idealized, wind tunnel experiments, atmospheric turbulence <span class="hlt">field</span> measurements, multi-year ABL tower <span class="hlt">observations</span>, and mesoscale models simulations, this study reveals remarkable similiarities (and some differences) between the small and larger scale components of the probability density functions increments <span class="hlt">fields</span>. This comprehensive analysis also utilizes a set of statistical distributions to showcase their ability to capture features of the <span class="hlt">velocity</span> and temperature increments' probability density functions (pdfs) across multiscale atmospheric motions. An approach is proposed for estimating their pdfs utilizing the maximum likelihood estimation (MLE) technique, which has never been conducted utilizing atmospheric data. Using this technique, we reveal the ability to estimate higher-order moments accurately with a limited sample size, which has been a persistent concern for atmospheric turbulence research. With the use robust Goodness of Fit (GoF) metrics, we quantitatively reveal the accuracy of the distributions to the diverse dataset. Through this analysis, it is shown that the normal inverse Gaussian (NIG) distribution is a prime candidate to be used as an estimate of the increment pdfs <span class="hlt">fields</span>. Therefore, using the NIG model and its parameters, we display the variations in the increments over a range of scales revealing some unique scale-dependent qualities under various stability and ow conditions. This</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AAS...23134009R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AAS...23134009R"><span>Dark Matter Profiles in Dwarf Galaxies: A Statistical Sample Using High-Resolution Hα <span class="hlt">Velocity</span> <span class="hlt">Fields</span> from PCWI</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Relatores, Nicole C.; Newman, Andrew B.; Simon, Joshua D.; Ellis, Richard; Truong, Phuongmai N.; Blitz, Leo</p> <p>2018-01-01</p> <p>We present high quality Hα <span class="hlt">velocity</span> <span class="hlt">fields</span> for a sample of nearby dwarf galaxies (log M/M⊙ = 8.4-9.8) obtained as part of the Dark Matter in Dwarf Galaxies survey. The purpose of the survey is to investigate the cusp-core discrepancy by quantifying the variation of the inner slope of the dark matter distributions of 26 dwarf galaxies, which were selected as likely to have regular kinematics. The data were obtained with the Palomar Cosmic Web Imager, located on the Hale 5m telescope. We extract rotation curves from the <span class="hlt">velocity</span> <span class="hlt">fields</span> and use optical and infrared photometry to model the stellar mass distribution. We model the total mass distribution as the sum of a generalized Navarro-Frenk-White dark matter halo along with the stellar and gaseous components. We present the distribution of inner dark matter density profile slopes derived from this analysis. For a subset of galaxies, we compare our results to an independent analysis based on CO <span class="hlt">observations</span>. In future work, we will compare the scatter in inner density slopes, as well as their correlations with galaxy properties, to theoretical predictions for dark matter core creation via supernovae feedback.</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('https://ntrs.nasa.gov/search.jsp?R=19830066192&hterms=seasonal+forecast&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dseasonal%2Bforecast','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830066192&hterms=seasonal+forecast&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dseasonal%2Bforecast"><span>Estimates of the seasonal mean vertical <span class="hlt">velocity</span> <span class="hlt">fields</span> of the extratropical Northern Hemisphere</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>White, G. H.</p> <p>1983-01-01</p> <p>Indirect methods are employed to estimate the wintertime and summertime mean vertical <span class="hlt">velocity</span> <span class="hlt">fields</span> of the extratropical Northern Hemisphere and intercomparisons are made, together with comparisons with mean seasonal patterns of cloudiness and precipitation. Twice-daily NMC operational analyses produced general circulation statistics for 11 winters and 12 summers, permitting calculation of the seasonal NMC averages for 6 hr forecasts, solution of the omega equation, integration of continuity equation downward from 100 mb, and solution of the thermodynamic energy equation in the absence of diabatic heating. The methods all yielded similar vertical <span class="hlt">velocity</span> patterns; however, the magnitude of the vertical <span class="hlt">velocities</span> could not be calculated with great accuracy. Orography was concluded to have less of an effect in summer than in winter, when winds are stronger.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15.9864P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15.9864P"><span>Concurrent <span class="hlt">field</span> measurements of turbulent <span class="hlt">velocities</span>, plant reconfiguration and drag forces on Ranunculus penicillatus</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Paul, Maike; Thomas, Robert E.; Keevil, Gareth M.</p> <p>2013-04-01</p> <p>-made drag sensor that was deployed flush with the streambed. Simultaneously, a profiling Acoustic Doppler Velocimeter (Nortek Vectrino-II) was deployed 0.5 m upstream of the plants. Also, a video camera was installed with its <span class="hlt">field</span> of view perpendicular to the mean flow direction, in order to record plant motion and reconfiguration associated with turbulent <span class="hlt">velocity</span> and drag fluctuations. Measurements were repeated while the Vectrino-II was consecutively deployed at four vertical positions to: 1. obtain a <span class="hlt">velocity</span> profile through the entire water column and 2. study which vertical position correlated most strongly to the drag force. <span class="hlt">Velocity</span> measurements confirmed that turbulent structures were present throughout the water column and a response to these fluctuations was <span class="hlt">observed</span> in the drag measurements. Responses lagged in time due to the horizontal distance between Vectrino-II and drag sensor position. Additionally, spectral analysis showed that the drag fluctuates with a frequency of 0.5 Hz which corresponds well with the undulating, quasi-sinusoidal, plant motion <span class="hlt">observed</span> on the video footage. This motion was associated with the downstream propagation of coherent eddies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AAS...20916501F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AAS...20916501F"><span>Using Open Clusters to Trace the Local Milky Way Rotation Curve and <span class="hlt">Velocity</span> <span class="hlt">Field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Frinchaboy, Peter M.; Majewski, S. R.</p> <p>2006-12-01</p> <p>Establishing the rotation curve of the Milky Way is one of the fundamental contributions needed to understand the Galaxy and its mass distribution. We have undertaken a systematic spectroscopic survey of open star clusters which can serve as tracers of Galactic disk dynamics. We report on our initial sample of 67 clusters for which the Hydra multi-fiber spectrographs on the WIYN and Blanco telescopes have delivered 1-2 km/s radial <span class="hlt">velocities</span> (RVs) of many dozens of stars in the <span class="hlt">fields</span> of each cluster, which are used to derive cluster membership and bulk cluster kinematics when combined with Tycho-2 proper motions. The clusters selected for study have a broad spatial distribution in order to be sensitive to the disk <span class="hlt">velocity</span> <span class="hlt">field</span> in all Galactic quadrants and across a Galactocentric radius range as much as 3.0 kpc from the solar circle. Through analysis of the cluster sample, we find (1) the rotation <span class="hlt">velocity</span> of the LSR is 221 (+2,-4) km/s, (2) the local rotation curve is declining with radius having a slope of -9.0 km/s/kpc, (3) we find (using R_0 = 8.5 kpc) the following Galactic parameters: A = 17.0 km/s/kpc and B = -8.9 km/s/kpc, which yields a Galaxy mass within of 1.5 R_0 of M = 0.9 ± 0.2 x 10^11 solar masses and a M/L of 5.9 in solar units. We also explore the distribution of the local <span class="hlt">velocity</span> <span class="hlt">field</span> and find evidence for non-circular motion due to the sprial arms.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013ExFl...54.1451S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ExFl...54.1451S"><span><span class="hlt">Velocity</span> measurements in the near <span class="hlt">field</span> of a diesel fuel injector by ultrafast imagery</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sedarsky, David; Idlahcen, Saïd; Rozé, Claude; Blaisot, Jean-Bernard</p> <p>2013-02-01</p> <p>This paper examines the <span class="hlt">velocity</span> profile of fuel issuing from a high-pressure single-orifice diesel injector. <span class="hlt">Velocities</span> of liquid structures were determined from time-resolved ultrafast shadow images, formed by an amplified two-pulse laser source coupled to a double-frame camera. A statistical analysis of the data over many injection events was undertaken to map <span class="hlt">velocities</span> related to spray formation near the nozzle outlet as a function of time after start of injection. These results reveal a strong asymmetry in the liquid profile of the test injector, with distinct fast and slow regions on opposite sides of the orifice. Differences of ˜100 m/s can be <span class="hlt">observed</span> between the `fast' and `slow' sides of the jet, resulting in different atomization conditions across the spray. On average, droplets are dispersed at a greater distance from the nozzle on the `fast' side of the flow, and distinct macrostructure can be <span class="hlt">observed</span> under the asymmetric <span class="hlt">velocity</span> conditions. The changes in structural <span class="hlt">velocity</span> and atomization behavior resemble flow structures which are often <span class="hlt">observed</span> in the presence of string cavitation produced under controlled conditions in scaled, transparent test nozzles. These <span class="hlt">observations</span> suggest that widely used common-rail supply configurations and modern injectors can potentially generate asymmetric interior flows which strongly influence diesel spray morphology. The velocimetry measurements presented in this work represent an effective and relatively straightforward approach to identify deviant flow behavior in real diesel sprays, providing new spatially resolved information on fluid structure and flow characteristics within the shear layers on the jet periphery.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29208976','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29208976"><span>Dynamical system with plastic self-organized <span class="hlt">velocity</span> <span class="hlt">field</span> as an alternative conceptual model of a cognitive system.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Janson, Natalia B; Marsden, Christopher J</p> <p>2017-12-05</p> <p>It is well known that architecturally the brain is a neural network, i.e. a collection of many relatively simple units coupled flexibly. However, it has been unclear how the possession of this architecture enables higher-level cognitive functions, which are unique to the brain. Here, we consider the brain from the viewpoint of dynamical systems theory and hypothesize that the unique feature of the brain, the self-organized plasticity of its architecture, could represent the means of enabling the self-organized plasticity of its <span class="hlt">velocity</span> vector <span class="hlt">field</span>. We propose that, conceptually, the principle of cognition could amount to the existence of appropriate rules governing self-organization of the <span class="hlt">velocity</span> <span class="hlt">field</span> of a dynamical system with an appropriate account of stimuli. To support this hypothesis, we propose a simple non-neuromorphic mathematical model with a plastic self-organized <span class="hlt">velocity</span> <span class="hlt">field</span>, which has no prototype in physical world. This system is shown to be capable of basic cognition, which is illustrated numerically and with musical data. Our conceptual model could provide an additional insight into the working principles of the brain. Moreover, hardware implementations of plastic <span class="hlt">velocity</span> <span class="hlt">fields</span> self-organizing according to various rules could pave the way to creating artificial intelligence of a novel type.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005AGUFM.G11A1203C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005AGUFM.G11A1203C"><span>Determining Sea-Level Rise and Coastal Subsidence in the Canadian Arctic Using a Dense GPS <span class="hlt">Velocity</span> <span class="hlt">Field</span> for North America</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Craymer, M.; Forbes, D.; Henton, J.; Lapelle, E.; Piraszewski, M.; Solomon, S.</p> <p>2005-12-01</p> <p>With <span class="hlt">observed</span> climate warming in the western Canadian Arctic and potential increases in regional sea level, we anticipate expansion of the coastal region subject to rising relative sea level and increased flooding risk. This is a concern for coastal communities such as Tuktoyaktuk and Sachs Harbour and for the design and safety of hydrocarbon production facilities on the Mackenzie Delta. To provide a framework in which to monitor these changes, a consistent <span class="hlt">velocity</span> <span class="hlt">field</span> has been determined from GPS <span class="hlt">observations</span> throughout North America, including the Canadian Arctic Archipelago and the Mackenzie Delta region. An expanded network of continuous GPS sites and multi-epoch (episodic) sites has enabled an increased density that enhances the application to geophysical studies including the discrimination of crustal motion, other components of coastal subsidence, and sea-level rise. To obtain a dense <span class="hlt">velocity</span> <span class="hlt">field</span> consistent at all scales, we have combined weekly solutions of continuous GPS sites from different agencies in Canada and the USA, together with the global reference frame under the North American Reference Frame initiative. Although there is already a high density of continuous GPS sites in the conterminous United States, there are many fewer such sites in Canada. To make up for this lack of density, we have incorporated high-accuracy episodic GPS <span class="hlt">observations</span> on stable monuments distributed throughout Canada. By combining up to ten years of repeated, episodic GPS <span class="hlt">observations</span> at such sites, together with weekly solutions from the continuous sites, we have obtained a highly consistent <span class="hlt">velocity</span> <span class="hlt">field</span> with a significantly increased spatial sampling of crustal deformation throughout Canada. This exhibits a spatially coherent pattern of uplift and subsidence in Canada that is consistent with the expected rates of glacial isostatic adjustment. To determine the contribution of vertical motion to sea-level rise under climate warming in the Canadian Arctic, we have</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PApGe.171..809B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PApGe.171..809B"><span>Modelling the <span class="hlt">Velocity</span> <span class="hlt">Field</span> in a Regular Grid in the Area of Poland on the Basis of the <span class="hlt">Velocities</span> of European Permanent Stations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bogusz, Janusz; Kłos, Anna; Grzempowski, Piotr; Kontny, Bernard</p> <p>2014-06-01</p> <p>The paper presents the results of testing the various methods of permanent stations' <span class="hlt">velocity</span> residua interpolation in a regular grid, which constitutes a continuous model of the <span class="hlt">velocity</span> <span class="hlt">field</span> in the territory of Poland. Three packages of software were used in the research from the point of view of interpolation: GMT ( The Generic Mapping Tools), Surfer and ArcGIS. The following methods were tested in the softwares: the Nearest Neighbor, Triangulation (TIN), Spline Interpolation, Surface, Inverse Distance to a Power, Minimum Curvature and Kriging. The presented research used the absolute <span class="hlt">velocities</span>' values expressed in the ITRF2005 reference frame and the intraplate <span class="hlt">velocities</span> related to the NUVEL model of over 300 permanent reference stations of the EPN and ASG-EUPOS networks covering the area of Europe. Interpolation for the area of Poland was done using data from the whole area of Europe to make the results at the borders of the interpolation area reliable. As a result of this research, an optimum method of such data interpolation was developed. All the mentioned methods were tested for being local or global, for the possibility to compute errors of the interpolated values, for explicitness and fidelity of the interpolation functions or the smoothing mode. In the authors' opinion, the best data interpolation method is Kriging with the linear semivariogram model run in the Surfer programme because it allows for the computation of errors in the interpolated values and it is a global method (it distorts the results in the least way). Alternately, it is acceptable to use the Minimum Curvature method. Empirical analysis of the interpolation results obtained by means of the two methods showed that the results are identical. The tests were conducted using the intraplate <span class="hlt">velocities</span> of the European sites. Statistics in the form of computing the minimum, maximum and mean values of the interpolated North and East components of the <span class="hlt">velocity</span> residuum were prepared for all</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/2018MeScT..29e5001H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MeScT..29e5001H"><span>Calculation of acoustic <span class="hlt">field</span> based on laser-measured vibration <span class="hlt">velocities</span> on ultrasonic transducer surface</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hu, Liang; Zhao, Nannan; Gao, Zhijian; Mao, Kai; Chen, Wenyu; Fu, Xin</p> <p>2018-05-01</p> <p>Determination of the distribution of a generated acoustic <span class="hlt">field</span> is valuable for studying ultrasonic transducers, including providing the guidance for transducer design and the basis for analyzing their performance, etc. A method calculating the acoustic <span class="hlt">field</span> based on laser-measured vibration <span class="hlt">velocities</span> on the ultrasonic transducer surface is proposed in this paper. Without knowing the inner structure of the transducer, the acoustic <span class="hlt">field</span> outside it can be calculated by solving the governing partial differential equation (PDE) of the <span class="hlt">field</span> based on the specified boundary conditions (BCs). In our study, the BC on the transducer surface, i.e. the distribution of the vibration <span class="hlt">velocity</span> on the surface, is accurately determined by laser scanning measurement of discrete points and follows a data fitting computation. In addition, to ensure the calculation accuracy for the whole <span class="hlt">field</span> even in an inhomogeneous medium, a finite element method is used to solve the governing PDE based on the mixed BCs, including the discretely measured <span class="hlt">velocity</span> data and other specified BCs. The method is firstly validated on numerical piezoelectric transducer models. The acoustic pressure distributions generated by a transducer operating in an homogeneous and inhomogeneous medium, respectively, are both calculated by the proposed method and compared with the results from other existing methods. Then, the method is further experimentally validated with two actual ultrasonic transducers used for flow measurement in our lab. The amplitude change of the output voltage signal from the receiver transducer due to changing the relative position of the two transducers is calculated by the proposed method and compared with the experimental data. This method can also provide the basis for complex multi-physical coupling computations where the effect of the acoustic <span class="hlt">field</span> should be taken into account.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.T11B0464T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.T11B0464T"><span>True-triaxial experimental seismic <span class="hlt">velocities</span> linked to an in situ 3D seismic <span class="hlt">velocity</span> structure</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tibbo, M.; Young, R. P.</p> <p>2017-12-01</p> <p>Upscaling from laboratory seismic <span class="hlt">velocities</span> to in situ <span class="hlt">field</span> seismic <span class="hlt">velocities</span> is a fundamental problem in rock physics. This study presents a unique situation where a 3D <span class="hlt">velocity</span> structure of comparable frequency ranges is available both in situ and experimentally. The in situ data comes from the Underground Research Laboratory (URL) located in Manitoba, Canada. The <span class="hlt">velocity</span> survey and oriented, cubic rock sample, are from the 420m level of the mine, where the geology is a homogeneous and isotropic granite. The triaxial in situ stress <span class="hlt">field</span> at this level was determined and the Mine-by tunnel was excavated horizontally to maximize borehole break out. Ultrasonic <span class="hlt">velocity</span> measurements for P-, S1-,and S2-waves were done in the tunnel sidewall, ceiling and far-<span class="hlt">field</span> rock mass.The geophysical imaging cell (GIC) used in this study allows for true triaxial stress (σ1 > σ2 > σ3). <span class="hlt">Velocity</span> surveys for P-, S1-, and S2-wave can be acquired along all three axes, and therefore the effects of σ1, σ2, σ3 on the <span class="hlt">velocity</span>-stress relationship is obtained along all 3 axes. The cubic (80 mm) granite sample was prepared oriented to the in situ principle stress axis in the <span class="hlt">field</span>. The stress path of the sample extraction from in situ stress was modeled in FLAC 3D (by Itasca inc ), and then reapplied in the GIC to obtain the laboratory <span class="hlt">velocities</span> at in situ stress. Both laboratory and <span class="hlt">field</span> <span class="hlt">velocities</span> conclude the same maximum <span class="hlt">velocity</span> axis, within error, to be along σ2 at 5880±60 m/s for P-wave. This deviation from the expected fast axis being σ1, is believed to be caused by an aligned microcrack fabric. The theory of acoustoelasticity, the dependence of acoustic wave <span class="hlt">velocity</span> on stresses in the propagating isotropic medium, is applied to the borehole hoop and radial stresses produced by the Mine-by tunnel. The acoustoelastic effect involves determining the linear (second-order) and nonlinear (third-order) elastic constants, which are derived from the <span class="hlt">velocity</span>-stress slopes</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> </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://adsabs.harvard.edu/abs/2012Icar..221..632B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012Icar..221..632B"><span><span class="hlt">Field</span> measurements of horizontal forward motion <span class="hlt">velocities</span> of terrestrial dust devils: Towards a proxy for ambient winds on Mars and Earth</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Balme, M. R.; Pathare, A.; Metzger, S. M.; Towner, M. C.; Lewis, S. R.; Spiga, A.; Fenton, L. K.; Renno, N. O.; Elliott, H. M.; Saca, F. A.; Michaels, T. I.; Russell, P.; Verdasca, J.</p> <p>2012-11-01</p> <p>Dust devils - convective vortices made visible by the dust and debris they entrain - are common in arid environments and have been <span class="hlt">observed</span> on Earth and Mars. Martian dust devils have been identified both in images taken at the surface and in remote sensing <span class="hlt">observations</span> from orbiting spacecraft. <span class="hlt">Observations</span> from landing craft and orbiting instruments have allowed the dust devil translational forward motion (ground <span class="hlt">velocity</span>) to be calculated, but it is unclear how these <span class="hlt">velocities</span> relate to the local ambient wind conditions, for (i) only model wind speeds are generally available for Mars, and (ii) on Earth only anecdotal evidence exists that compares dust devil ground <span class="hlt">velocity</span> with ambient wind <span class="hlt">velocity</span>. If dust devil ground <span class="hlt">velocity</span> can be reliably correlated to the ambient wind regime, <span class="hlt">observations</span> of dust devils could provide a proxy for wind speed and direction measurements on Mars. Hence, dust devil ground <span class="hlt">velocities</span> could be used to probe the circulation of the martian boundary layer and help constrain climate models or assess the safety of future landing sites. We present results from a <span class="hlt">field</span> study of terrestrial dust devils performed in the southwest USA in which we measured dust devil horizontal <span class="hlt">velocity</span> as a function of ambient wind <span class="hlt">velocity</span>. We acquired stereo images of more than a 100 active dust devils and recorded multiple size and position measurements for each dust devil. We used these data to calculate dust devil translational <span class="hlt">velocity</span>. The dust devils were within a study area bounded by 10 m high meteorology towers such that dust devil speed and direction could be correlated with the local ambient wind speed and direction measurements. Daily (10:00-16:00 local time) and 2-h averaged dust devil ground speeds correlate well with ambient wind speeds averaged over the same period. Unsurprisingly, individual measurements of dust devil ground speed match instantaneous measurements of ambient wind speed more poorly; a 20-min smoothing window applied to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009Th%26Ae..16..325A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009Th%26Ae..16..325A"><span>Application of ``POLIS'' PIV system for measurement of <span class="hlt">velocity</span> <span class="hlt">fields</span> in a supersonic flow of the wind tunnels</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Akhmetbekov, Y. K.; Bilsky, A. V.; Markovich, D. M.; Maslov, A. A.; Polivanov, P. A.; Tsyryul'Nikov, I. S.; Yaroslavtsev, M. I.</p> <p>2009-09-01</p> <p>Measurement results on the mean <span class="hlt">velocity</span> <span class="hlt">fields</span> and <span class="hlt">fields</span> of <span class="hlt">velocity</span> pulsations in the supersonic flows obtained by means of the PIV measurement set “POLIS” are presented. Experiments were carried out in the supersonic blow-down and stationary wind tunnels at the Mach numbers of 4.85 and 6. The method of flow <span class="hlt">velocity</span> estimate in the test section of the blow-down wind tunnel was grounded by direct measurements of stagnation pressure in the setup settling chamber. The size of tracer particles introduced into the supersonic flow by a mist generator was determined; data on the structure of pulsating <span class="hlt">velocity</span> in a track of an oblique-cut gas-dynamic whistle were obtained under the conditions of self-oscillations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhDT........46S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhDT........46S"><span>Ramifications of projectile <span class="hlt">velocity</span> on the ballistic dart penetration of sand</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sable, Peter Anthony</p> <p></p> <p>With the advent of novel in-situ experimental measurement techniques, highly resolved quantitative <span class="hlt">observations</span> of dynamic events within granular media can now be made. In particular, high speed imagery and digital analysis now allow for the ballistic behaviors of sand to be examined not only across a range of event <span class="hlt">velocities</span> but across multiple length scales. In an attempt to further understand the dynamic behavior of granular media, these new experimental developments were implemented utilizing high speed photography coupled with piezo-electric stress gauges to <span class="hlt">observe</span> visually accessible ballistic events of a dart penetrating Ottawa sand. Projectile <span class="hlt">velocities</span> ranged from 100 to over 300 meters per second with two distinct chosen <span class="hlt">fields</span> of view to capture bulk and grain-scale behaviors. Each event was analyzed using the digital image correlation technique, particle image velocimetry from which two dimensional, temporally resolved, <span class="hlt">velocity</span> <span class="hlt">fields</span> were extracted, from which bulk granular flow and compaction wave propagation were <span class="hlt">observed</span> and quantified. By comparing bulk, in situ, <span class="hlt">velocity</span> <span class="hlt">field</span> behavior resultant from dart penetration, momentum transfer could be quantified measuring radius of influence or dilatant fluid approximations from which a positive correlation was found across the explored <span class="hlt">velocity</span> regime, including self similar tendencies. This was, however, not absolute as persistent scatter was <span class="hlt">observed</span> attributed to granular heterogeneous effects. These were tentatively measured in terms of an irreversible energy amount calculated via energy balance. Grain scale analysis reveals analogous behavior to the bulk response with more chaotic structure, though conclusions were limited by the image processing method to qualitative <span class="hlt">observations</span>. Even so, critical granular behaviors could be seen, such as densification, pore collapse, and grain fracture from which basic heterogeneous phenomena could be examined. These particularly dominated near nose</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19880005580','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19880005580"><span>Lift distribution and <span class="hlt">velocity</span> <span class="hlt">field</span> measurements for a three-dimensional, steady blade/vortex interaction</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Dunagan, Stephen E.; Norman, Thomas R.</p> <p>1987-01-01</p> <p>A wind tunnel experiment simulating a steady three-dimensional helicopter rotor blade/vortex interaction is reported. The experimental configuration consisted of a vertical semispan vortex-generating wing, mounted upstream of a horizontal semispan rotor blade airfoil. A three-dimensional laser velocimeter was used to measure the <span class="hlt">velocity</span> <span class="hlt">field</span> in the region of the blade. Sectional lift coefficients were calculated by integrating the <span class="hlt">velocity</span> <span class="hlt">field</span> to obtain the bound vorticity. Total lift values, obtained by using an internal strain-gauge balance, verified the laser velocimeter data. Parametric variations of vortex strength, rotor blade angle of attack, and vortex position relative to the rotor blade were explored. These data are reported (with attention to experimental limitations) to provide a dataset for the validation of analytical work.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003APS..DFD.FE008K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003APS..DFD.FE008K"><span><span class="hlt">Velocity</span> <span class="hlt">Field</span> Measurements of Human Coughing Using Time Resolved Particle Image Velocimetry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Khan, T.; Marr, D. R.; Higuchi, H.; Glauser, M. N.</p> <p>2003-11-01</p> <p>Quantitative fluid mechanics analysis of human coughing has been carried out using new Time Resolved Particle Image Velocimetry (TRPIV). The study involves measurement of <span class="hlt">velocity</span> vector time-histories and <span class="hlt">velocity</span> profiles. It is focused on the average normal human coughing. Some work in the past on cough mechanics has involved measurement of flow rates, tidal volumes and sub-glottis pressure. However, data of unsteady <span class="hlt">velocity</span> vector <span class="hlt">field</span> of the exiting highly time-dependent jets is not available. In this study, human cough waveform data are first acquired in vivo using conventional respiratory instrumentation for various volunteers of different gender/age groups. The representative waveform is then reproduced with a coughing/breathing simulator (with or without a manikin) for TRPIV measurements and analysis. The results of this study would be useful not only for designing of indoor air quality and heating, ventilation and air conditioning systems, but also for devising means of protection against infectious diseases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21266556','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21266556"><span>Flow <span class="hlt">velocity</span> vector <span class="hlt">fields</span> by ultrasound particle imaging velocimetry: in vitro comparison with optical flow velocimetry.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Westerdale, John; Belohlavek, Marek; McMahon, Eileen M; Jiamsripong, Panupong; Heys, Jeffrey J; Milano, Michele</p> <p>2011-02-01</p> <p>We performed an in vitro study to assess the precision and accuracy of particle imaging velocimetry (PIV) data acquired using a clinically available portable ultrasound system via comparison with stereo optical PIV. The performance of ultrasound PIV was compared with optical PIV on a benchmark problem involving vortical flow with a substantial out-of-plane <span class="hlt">velocity</span> component. Optical PIV is capable of stereo image acquisition, thus measuring out-of-plane <span class="hlt">velocity</span> components. This allowed us to quantify the accuracy of ultrasound PIV, which is limited to in-plane acquisition. The system performance was assessed by considering the instantaneous <span class="hlt">velocity</span> <span class="hlt">fields</span> without extracting <span class="hlt">velocity</span> profiles by spatial averaging. Within the 2-dimensional correlation window, using 7 time-averaged frames, the vector <span class="hlt">fields</span> were found to have correlations of 0.867 in the direction along the ultrasound beam and 0.738 in the perpendicular direction. Out-of-plane motion of greater than 20% of the in-plane vector magnitude was found to increase the SD by 11% for the vectors parallel to the ultrasound beam direction and 8.6% for the vectors perpendicular to the beam. The results show a close correlation and agreement of individual <span class="hlt">velocity</span> vectors generated by ultrasound PIV compared with optical PIV. Most of the measurement distortions were caused by out-of-plane <span class="hlt">velocity</span> components.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.S31E..08E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.S31E..08E"><span>Rapid Non-Gaussian Uncertainty Quantification of Seismic <span class="hlt">Velocity</span> Models and Images</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ely, G.; Malcolm, A. E.; Poliannikov, O. V.</p> <p>2017-12-01</p> <p>Conventional seismic imaging typically provides a single estimate of the subsurface without any error bounds. Noise in the <span class="hlt">observed</span> raw traces as well as the uncertainty of the <span class="hlt">velocity</span> model directly impact the uncertainty of the final seismic image and its resulting interpretation. We present a Bayesian inference framework to quantify uncertainty in both the <span class="hlt">velocity</span> model and seismic images, given noise statistics of the <span class="hlt">observed</span> data.To estimate <span class="hlt">velocity</span> model uncertainty, we combine the <span class="hlt">field</span> expansion method, a fast frequency domain wave equation solver, with the adaptive Metropolis-Hastings algorithm. The speed of the <span class="hlt">field</span> expansion method and its reduced parameterization allows us to perform the tens or hundreds of thousands of forward solves needed for non-parametric posterior estimations. We then migrate the <span class="hlt">observed</span> data with the distribution of <span class="hlt">velocity</span> models to generate uncertainty estimates of the resulting subsurface image. This procedure allows us to create both qualitative descriptions of seismic image uncertainty and put error bounds on quantities of interest such as the dip angle of a subduction slab or thickness of a stratigraphic layer.</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 <span class="hlt">field</span> 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 <span class="hlt">field</span> 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> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUOSPO21A..06B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUOSPO21A..06B"><span>Inference and Biogeochemical Response of Vertical <span class="hlt">Velocities</span> inside a Mode Water Eddy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barceló-Llull, B.; Pallas Sanz, E.; Sangrà, P.</p> <p>2016-02-01</p> <p>With the aim to study the modulation of the biogeochemical fluxes by the ageostrophic secondary circulation in anticyclonic mesoscale eddies, a typical eddy of the Canary Eddy Corridor was interdisciplinary surveyed on September 2014 in the framework of the PUMP project. The eddy was elliptical shaped, 4 month old, 110 km diameter and 400 m depth. It was an intrathermocline type often also referred as mode water eddy type. We inferred the mesoscale vertical <span class="hlt">velocity</span> <span class="hlt">field</span> resolving a generalized omega equation from the 3D density and ADCP <span class="hlt">velocity</span> <span class="hlt">fields</span> of a five-day sampled CTD-SeaSoar regular grid centred on the eddy. The grid transects where 10 nautical miles apart. Although complex, in average, the inferred omega <span class="hlt">velocity</span> <span class="hlt">field</span> (hereafter w) shows a dipolar structure with downwelling <span class="hlt">velocities</span> upstream of the propagation path (west) and upwelling <span class="hlt">velocities</span> downstream. The w at the eddy center was zero and maximum values were located at the periphery attaining ca. 6 m day-1. Coinciding with the occurrence of the vertical <span class="hlt">velocities</span> cells a noticeable enhancement of phytoplankton biomass was <span class="hlt">observed</span> at the eddy periphery respect to the far <span class="hlt">field</span>. A corresponding upward diapycnal flux of nutrients was also <span class="hlt">observed</span> at the periphery. As minimum <span class="hlt">velocities</span> where reached at the eddy center, lineal Ekman pumping mechanism was discarded. Minimum values of phytoplankton biomass where also <span class="hlt">observed</span> at the eddy center. The possible mechanisms for such dipolar w cell are still being investigated, but an analysis of the generalized omega equation forcing terms suggest that it may be a combination of horizontal deformation and advection of vorticity by the ageostrophic current (related to nonlinear Ekman pumping). As expected for Trades, the wind was rather constant and uniform with a speed of ca. 5 m s-1. Diagnosed nonlinear Ekman pumping leaded also to a dipolar cell that mirrors the omega w dipolar cell.</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><span class="hlt">Field</span> 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 <span class="hlt">field</span> 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 gradient along the flow direction, remotely sensed estimates of discharge appear to be viable.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20070030267','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20070030267"><span><span class="hlt">Velocity-Field</span> Measurements of an Axisymmetric Separated Flow Subjected to Amplitude-Modulated Excitation</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Trosin, Barry James</p> <p>2007-01-01</p> <p>Active flow control was applied at the point of separation of an axisymmetric, backward-facing-step flow. The control was implemented by employing a Helmholtz resonator that was externally driven by an amplitude-modulated, acoustic disturbance from a speaker located upstream of the wind tunnel. The <span class="hlt">velocity</span> <span class="hlt">field</span> of the separating/reattaching flow region downstream of the step was characterized using hotwire <span class="hlt">velocity</span> measurements with and without flow control. Conventional statistics of the data reveal that the separating/reattaching flow is affected by the imposed forcing. Triple decomposition along with conditional averaging was used to distinguish periodic disturbances from random turbulence in the fluctuating <span class="hlt">velocity</span> component. A significant outcome of the present study is that it demonstrates that amplitude-modulated forcing of the separated flow alters the flow in the same manner as the more conventional method of periodic excitation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018HydJ..tmp...86W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018HydJ..tmp...86W"><span>The impact of groundwater <span class="hlt">velocity</span> <span class="hlt">fields</span> on streamlines in an aquifer system with a discontinuous aquitard (Inner Mongolia, China)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Qiang; Zhao, Yingwang; Xu, Hua</p> <p>2018-04-01</p> <p>Many numerical methods that simulate groundwater flow, particularly the continuous Galerkin finite element method, do not produce <span class="hlt">velocity</span> information directly. Many algorithms have been proposed to improve the accuracy of <span class="hlt">velocity</span> <span class="hlt">fields</span> computed from hydraulic potentials. The differences in the streamlines generated from <span class="hlt">velocity</span> <span class="hlt">fields</span> obtained using different algorithms are presented in this report. The superconvergence method employed by FEFLOW, a popular commercial code, and some dual-mesh methods proposed in recent years are selected for comparison. The applications to depict hydrogeologic conditions using streamlines are used, and errors in streamlines are shown to lead to notable errors in boundary conditions, the locations of material interfaces, fluxes and conductivities. Furthermore, the effects of the procedures used in these two types of methods, including <span class="hlt">velocity</span> integration and local conservation, are analyzed. The method of interpolating <span class="hlt">velocities</span> across edges using fluxes is shown to be able to eliminate errors associated with refraction points that are not located along material interfaces and streamline ends at no-flow boundaries. Local conservation is shown to be a crucial property of <span class="hlt">velocity</span> <span class="hlt">fields</span> and can result in more accurate streamline densities. A case study involving both three-dimensional and two-dimensional cross-sectional models of a coal mine in Inner Mongolia, China, are used to support the conclusions presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29289237','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29289237"><span>A <span class="hlt">velocity</span> probe-based method for continuous detonation and shock measurement in near-<span class="hlt">field</span> underwater explosion.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Kebin; Li, Xiaojie; Yan, Honghao; Wang, Xiaohong; Miao, Yusong</p> <p>2017-12-01</p> <p>A new <span class="hlt">velocity</span> probe which permits recording the time history of detonation and shock waves has been developed by improving the commercial on principle and structure. A method based on the probe is then designed to measure the detonation <span class="hlt">velocity</span> and near-<span class="hlt">field</span> shock parameters in a single underwater explosion, by which the oblique shock wave front of cylindrical charges and the peak pressure attenuation curve of spherical explosive are obtained. A further derivation of detonation pressure, adiabatic exponent, and other shock parameters is conducted. The present method offers a novel and reliable parameter determination for near-<span class="hlt">field</span> underwater explosion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017RScI...88l3905L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017RScI...88l3905L"><span>A <span class="hlt">velocity</span> probe-based method for continuous detonation and shock measurement in near-<span class="hlt">field</span> underwater explosion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Kebin; Li, Xiaojie; Yan, Honghao; Wang, Xiaohong; Miao, Yusong</p> <p>2017-12-01</p> <p>A new <span class="hlt">velocity</span> probe which permits recording the time history of detonation and shock waves has been developed by improving the commercial on principle and structure. A method based on the probe is then designed to measure the detonation <span class="hlt">velocity</span> and near-<span class="hlt">field</span> shock parameters in a single underwater explosion, by which the oblique shock wave front of cylindrical charges and the peak pressure attenuation curve of spherical explosive are obtained. A further derivation of detonation pressure, adiabatic exponent, and other shock parameters is conducted. The present method offers a novel and reliable parameter determination for near-<span class="hlt">field</span> underwater explosion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15697718','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15697718"><span>Anomalous scaling of a passive scalar advected by the Navier-Stokes <span class="hlt">velocity</span> <span class="hlt">field</span>: two-loop approximation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Adzhemyan, L Ts; Antonov, N V; Honkonen, J; Kim, T L</p> <p>2005-01-01</p> <p>The <span class="hlt">field</span> theoretic renormalization group and operator-product expansion are applied to the model of a passive scalar quantity advected by a non-Gaussian <span class="hlt">velocity</span> <span class="hlt">field</span> with finite correlation time. The <span class="hlt">velocity</span> is governed by the Navier-Stokes equation, subject to an external random stirring force with the correlation function proportional to delta(t- t')k(4-d-2epsilon). It is shown that the scalar <span class="hlt">field</span> is intermittent already for small epsilon, its structure functions display anomalous scaling behavior, and the corresponding exponents can be systematically calculated as series in epsilon. The practical calculation is accomplished to order epsilon2 (two-loop approximation), including anisotropic sectors. As for the well-known Kraichnan rapid-change model, the anomalous scaling results from the existence in the model of composite <span class="hlt">fields</span> (operators) with negative scaling dimensions, identified with the anomalous exponents. Thus the mechanism of the origin of anomalous scaling appears similar for the Gaussian model with zero correlation time and the non-Gaussian model with finite correlation time. It should be emphasized that, in contrast to Gaussian <span class="hlt">velocity</span> ensembles with finite correlation time, the model and the perturbation theory discussed here are manifestly Galilean covariant. The relevance of these results for real passive advection and comparison with the Gaussian models and experiments are briefly discussed.</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://www.ncbi.nlm.nih.gov/pubmed/29714472','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29714472"><span>Large-<span class="hlt">Velocity</span> Saturation in Thin-Film Black Phosphorus Transistors.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chen, Xiaolong; Chen, Chen; Levi, Adi; Houben, Lothar; Deng, Bingchen; Yuan, Shaofan; Ma, Chao; Watanabe, Kenji; Taniguchi, Takashi; Naveh, Doron; Du, Xu; Xia, Fengnian</p> <p>2018-05-22</p> <p>A high saturation <span class="hlt">velocity</span> semiconductor is appealing for applications in electronics and optoelectronics. Thin-film black phosphorus (BP), an emerging layered semiconductor, shows a high carrier mobility and strong mid-infrared photoresponse at room temperature. Here, we report the <span class="hlt">observation</span> of high intrinsic saturation <span class="hlt">velocity</span> in 7 to 11 nm thick BP for both electrons and holes as a function of charge-carrier density, temperature, and crystalline direction. We distinguish a drift <span class="hlt">velocity</span> transition point due to the competition between the electron-impurity and electron-phonon scatterings. We further achieve a room-temperature saturation <span class="hlt">velocity</span> of 1.2 (1.0) × 10 7 cm s -1 for hole (electron) carriers at a critical electric <span class="hlt">field</span> of 14 (13) kV cm -1 , indicating an intrinsic current-gain cutoff frequency ∼20 GHz·μm for radio frequency applications. Moreover, the current density is as high as 580 μA μm -1 at a low electric <span class="hlt">field</span> of 10 kV cm -1 . Our studies demonstrate that thin-film BP outperforms silicon in terms of saturation <span class="hlt">velocity</span> and critical <span class="hlt">field</span>, revealing its great potential in radio-frequency electronics, high-speed mid-infrared photodetectors, and optical modulators.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhCS.891a2037S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhCS.891a2037S"><span>Numerical simulation of <span class="hlt">velocity</span> and temperature <span class="hlt">fields</span> in natural circulation loop</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sukomel, L. A.; Kaban'kov, O. N.</p> <p>2017-11-01</p> <p>Low flow natural circulation regimes are realized in many practical applications and the existence of the reliable engineering and design calculation methods of flows driven exclusively by buoyancy forces is an actual problem. In particular it is important for the analysis of start up regimes of passive safety systems of nuclear power plants. In spite of a long year investigations of natural circulation loops no suitable predicting recommendations for heat transfer and friction for the above regimes have been proposed for engineering practice and correlations for forced flow are commonly used which considerably overpredicts the real flow <span class="hlt">velocities</span>. The 2D numerical simulation of <span class="hlt">velocity</span> and temperature <span class="hlt">fields</span> in circular tubes for laminar flow natural circulation with reference to the laboratory experimental loop has been carried out. The results were compared with the 1D modified model and experimental data obtained on the above loop. The 1D modified model was still based on forced flow correlations, but in these correlations the physical properties variability and the existence of thermal and hydrodynamic entrance regions are taken into account. The comparison of 2D simulation, 1D model calculations and the experimental data showed that even subject to influence of liquid properties variability and entrance regions on heat transfer and friction the use of 1D model with forced flow correlations do not improve the accuracy of calculations. In general, according to 2D numerical simulation the wall shear stresses are mainly affected by the change of wall <span class="hlt">velocity</span> gradient due to practically continuous <span class="hlt">velocity</span> profiles deformation along the whole heated zone. The form of <span class="hlt">velocity</span> profiles and the extent of their deformation in its turn depend upon the wall heat flux density and the hydraulic diameter.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015OcMod..89....1H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015OcMod..89....1H"><span>Optimizing <span class="hlt">velocities</span> and transports for complex coastal regions and archipelagos</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Haley, Patrick J.; Agarwal, Arpit; Lermusiaux, Pierre F. J.</p> <p>2015-05-01</p> <p>We derive and apply a methodology for the initialization of <span class="hlt">velocity</span> and transport <span class="hlt">fields</span> in complex multiply-connected regions with multiscale dynamics. The result is initial <span class="hlt">fields</span> that are consistent with <span class="hlt">observations</span>, complex geometry and dynamics, and that can simulate the evolution of ocean processes without large spurious initial transients. A class of constrained weighted least squares optimizations is defined to best fit first-guess <span class="hlt">velocities</span> while satisfying the complex bathymetry, coastline and divergence strong constraints. A weak constraint towards the minimum inter-island transports that are in accord with the first-guess <span class="hlt">velocities</span> provides important <span class="hlt">velocity</span> corrections in complex archipelagos. In the optimization weights, the minimum distance and vertical area between pairs of coasts are computed using a Fast Marching Method. Additional information on <span class="hlt">velocity</span> and transports are included as strong or weak constraints. We apply our methodology around the Hawaiian islands of Kauai/Niihau, in the Taiwan/Kuroshio region and in the Philippines Archipelago. Comparisons with other common initialization strategies, among hindcasts from these initial conditions (ICs), and with independent in situ <span class="hlt">observations</span> show that our optimization corrects transports, satisfies boundary conditions and redirects currents. Differences between the hindcasts from these different ICs are found to grow for at least 2-3 weeks. When compared to independent in situ <span class="hlt">observations</span>, simulations from our optimized ICs are shown to have the smallest errors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010TCD.....4.2415M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010TCD.....4.2415M"><span>Photogrammetric determination of spatio-temporal <span class="hlt">velocity</span> <span class="hlt">fields</span> at Glaciar San Rafael in the Northern Patagonian Icefield</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Maas, H.-G.; Casassa, G.; Schneider, D.; Schwalbe, E.; Wendt, A.</p> <p>2010-11-01</p> <p>Glaciar San Rafael in the Northern Patagonian Icefield, with a length of 46 km and an ice area of 722 km2, is the lowest latitude tidewater outlet glacier in the world and one of the fastest and most productive glaciers in southern South America in terms of iceberg flux. In a joint project of the TU Dresden and CECS, spatio-temporal <span class="hlt">velocity</span> <span class="hlt">fields</span> in the region of the glacier front were determined in a campaign in austral spring of 2009. Monoscopic terrestrial image sequences were recorded with an intervallometer mode high resolution digital camera over several days. In these image sequences, a large number of glacier surface points were tracked by subpixel accuracy feature tracking techniques. Scaling and georeferencing of the trajectories obtained from image space tracking was performed via a multi-station GPS-supported photogrammetric network. The technique allows for tracking hundreds of glacier surface points at a measurement accuracy in the order of one decimeter and an almost arbitrarily high temporary resolution. The results show <span class="hlt">velocities</span> of up to 16 m per day. No significant tidal signals could be <span class="hlt">observed</span>. Our <span class="hlt">velocities</span> are in agreement with earlier measurements from theodolite and satellite interferometry performed in 1986-1994, suggesting that the current thinning of 3.5 m/y at the front is not due to dynamic thinning but rather by enhanced melting.</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/2016ASPC..504..119S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ASPC..504..119S"><span>Magnetic <span class="hlt">Field</span> and Plasma Diagnostics from Coordinated Prominence <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>Schmieder, B.; Levens, P.; Dalmasse, K.; Mein, N.; Mein, P.; Lopez-Ariste, A.; Labrosse, N.; Heinzel, P.</p> <p>2016-04-01</p> <p>We study the magnetic <span class="hlt">field</span> in prominences from a statistical point of view, by using THEMIS in the MTR mode, performing spectropolarimetry of the He I D3 line. Combining these measurements with spectroscopic data from IRIS, Hinode/EIS as well as ground-based telescopes, such as the Meudon Solar Tower, we infer the temperature, density, and flow <span class="hlt">velocities</span> of the plasma. There are a number of open questions that we aim to answer: - What is the general direction of the magnetic <span class="hlt">field</span> in prominences? Is the model using a single orientation of magnetic <span class="hlt">field</span> always valid for atypical prominences? %- Does this depend on the location of the filament on the disk (visible in Hα, in He II 304 Å) over an inversion line between weak or strong network ? - Are prominences in a weak environment <span class="hlt">field</span> dominated by gas pressure? - Measuring the Doppler shifts in Mg II lines (with IRIS) and in Hα can tell us if there are substantial <span class="hlt">velocities</span> to maintain vertical rotating structures, as has been suggested for tornado-like prominences. We present here some results obtained with different ground-based and space-based instruments in this framework.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910055777&hterms=Quasi+experiment&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DQuasi%2Bexperiment','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910055777&hterms=Quasi+experiment&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DQuasi%2Bexperiment"><span>Interpretation of the electric <span class="hlt">fields</span> measured in an ionospheric 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>Brenning, N.; Faelthammar, C.-G.; Marklund, G.; Haerendel, G.; Kelley, M. C.; Pfaff, R.</p> <p>1991-01-01</p> <p>The quasi-dc electric <span class="hlt">fields</span> measured in the CRIT I ionospheric release experiment are studied. In the experiment, two identical barium shaped charges were fired toward a main payload, and three-dimensional measurements of the electric <span class="hlt">field</span> inside the streams were made. The relevance of proposed mechanisms for electron heating in the critical ionization <span class="hlt">velocity</span> (CIV) mechanism is addressed. It is concluded that both the 'homogeneous' and the 'ionizing front' models probably are valid, but in different parts of the streams. It is also possible that electrons are directly accelerated by a magnetic <span class="hlt">field</span>-aligned component of the electric <span class="hlt">field</span>. The coupling between the ambient ionosphere and the ionized barium stream is more complicated that is usually assumed in CIV theories, with strong magnetic-<span class="hlt">field</span>-aligned electric <span class="hlt">fields</span> and probably current limitation as important processes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22270624-magnetized-gas-smith-high-velocity-cloud','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22270624-magnetized-gas-smith-high-velocity-cloud"><span>MAGNETIZED GAS IN THE SMITH HIGH <span class="hlt">VELOCITY</span> CLOUD</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>Hill, Alex S.; McClure-Griffiths, Naomi M.; Mao, S. A.</p> <p>2013-11-01</p> <p>We report the first detection of magnetic <span class="hlt">fields</span> associated with the Smith High <span class="hlt">Velocity</span> Cloud. We use a catalog of Faraday rotation measures toward extragalactic radio sources behind the Smith Cloud, new H I <span class="hlt">observations</span> from the Robert C. Byrd Green Bank Telescope, and a spectroscopic map of Hα from the Wisconsin H-Alpha Mapper Northern Sky Survey. There are enhancements in rotation measure (RM) of ≈100 rad m{sup –2} which are generally well correlated with decelerated Hα emission. We estimate a lower limit on the line-of-sight component of the <span class="hlt">field</span> of ≈8 μG along a decelerated filament; this is amore » lower limit due to our assumptions about the geometry. No RM excess is evident in sightlines dominated by H I or Hα at the <span class="hlt">velocity</span> of the Smith Cloud. The smooth Hα morphology of the emission at the Smith Cloud <span class="hlt">velocity</span> suggests photoionization by the Galactic ionizing radiation <span class="hlt">field</span> as the dominant ionization mechanism, while the filamentary morphology and high (≈1 Rayleigh) Hα intensity of the lower-<span class="hlt">velocity</span> magnetized ionized gas suggests an ionization process associated with shocks due to interaction with the Galactic interstellar medium. The presence of the magnetic <span class="hlt">field</span> may contribute to the survival of high <span class="hlt">velocity</span> clouds like the Smith Cloud as they move from the Galactic halo to the disk. We expect these data to provide a test for magnetohydrodynamic simulations of infalling gas.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.475.4809S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.475.4809S"><span>Constraining cosmology with the <span class="hlt">velocity</span> function of low-mass galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schneider, Aurel; Trujillo-Gomez, Sebastian</p> <p>2018-04-01</p> <p>The number density of <span class="hlt">field</span> galaxies per rotation <span class="hlt">velocity</span>, referred to as the <span class="hlt">velocity</span> function, is an intriguing statistical measure probing the smallest scales of structure formation. In this paper we point out that the <span class="hlt">velocity</span> function is sensitive to small shifts in key cosmological parameters such as the amplitude of primordial perturbations (σ8) or the total matter density (Ωm). Using current data and applying conservative assumptions about baryonic effects, we show that the <span class="hlt">observed</span> <span class="hlt">velocity</span> function of the Local Volume favours cosmologies in tension with the measurements from Planck but in agreement with the latest findings from weak lensing surveys. While the current systematics regarding the relation between <span class="hlt">observed</span> and true rotation <span class="hlt">velocities</span> are potentially important, upcoming data from H I surveys as well as new insights from hydrodynamical simulations will dramatically improve the situation in the near future.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007MNRAS.375..348C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007MNRAS.375..348C"><span>Cosmic <span class="hlt">velocity</span>-gravity relation in redshift space</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Colombi, Stéphane; Chodorowski, Michał J.; Teyssier, Romain</p> <p>2007-02-01</p> <p>We propose a simple way to estimate the parameter β ~= Ω0.6/b from 3D galaxy surveys, where Ω is the non-relativistic matter-density parameter of the Universe and b is the bias between the galaxy distribution and the total matter distribution. Our method consists in measuring the relation between the cosmological <span class="hlt">velocity</span> and gravity <span class="hlt">fields</span>, and thus requires peculiar <span class="hlt">velocity</span> measurements. The relation is measured directly in redshift space, so there is no need to reconstruct the density <span class="hlt">field</span> in real space. In linear theory, the radial components of the gravity and <span class="hlt">velocity</span> <span class="hlt">fields</span> in redshift space are expected to be tightly correlated, with a slope given, in the distant <span class="hlt">observer</span> approximation, by We test extensively this relation using controlled numerical experiments based on a cosmological N-body simulation. To perform the measurements, we propose a new and rather simple adaptive interpolation scheme to estimate the <span class="hlt">velocity</span> and the gravity <span class="hlt">field</span> on a grid. One of the most striking results is that non-linear effects, including `fingers of God', affect mainly the tails of the joint probability distribution function (PDF) of the <span class="hlt">velocity</span> and gravity <span class="hlt">field</span>: the 1-1.5 σ region around the maximum of the PDF is dominated by the linear theory regime, both in real and redshift space. This is understood explicitly by using the spherical collapse model as a proxy of non-linear dynamics. Applications of the method to real galaxy catalogues are discussed, including a preliminary investigation on homogeneous (volume-limited) `galaxy' samples extracted from the simulation with simple prescriptions based on halo and substructure identification, to quantify the effects of the bias between the galaxy distribution and the total matter distribution, as well as the effects of shot noise.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016Icar..271...57C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016Icar..271...57C"><span>Characterizing the original ejection <span class="hlt">velocity</span> <span class="hlt">field</span> of the Koronis family</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Carruba, V.; Nesvorný, D.; Aljbaae, S.</p> <p>2016-06-01</p> <p>An asteroid family forms as a result of a collision between an impactor and a parent body. The fragments with ejection speeds higher than the escape <span class="hlt">velocity</span> from the parent body can escape its gravitational pull. The cloud of escaping debris can be identified by the proximity of orbits in proper element, or frequency, domains. Obtaining estimates of the original ejection speed can provide valuable constraints on the physical processes occurring during collision, and used to calibrate impact simulations. Unfortunately, proper elements of asteroids families are modified by gravitational and non-gravitational effects, such as resonant dynamics, encounters with massive bodies, and the Yarkovsky effect, such that information on the original ejection speeds is often lost, especially for older, more evolved families. It has been recently suggested that the distribution in proper inclination of the Koronis family may have not been significantly perturbed by local dynamics, and that information on the component of the ejection <span class="hlt">velocity</span> that is perpendicular to the orbital plane (vW), may still be available, at least in part. In this work we estimate the magnitude of the original ejection <span class="hlt">velocity</span> speeds of Koronis members using the <span class="hlt">observed</span> distribution in proper eccentricity and inclination, and accounting for the spread caused by dynamical effects. Our results show that (i) the spread in the original ejection speeds is, to within a 15% error, inversely proportional to the fragment size, and (ii) the minimum ejection <span class="hlt">velocity</span> is of the order of 50 m/s, with larger values possible depending on the orbital configuration at the break-up.</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> <span class="hlt">fields</span> 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> <span class="hlt">fields</span> 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/2015AGUFMSH41D2401T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMSH41D2401T"><span>Systematic Variability of the He+ Pickup Ion <span class="hlt">Velocity</span> Distribution Function <span class="hlt">Observed</span> with SOHO/CELIAS/CTOF</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.; Wimmer-Schweingruber, R. F.</p> <p>2015-12-01</p> <p>The 1D <span class="hlt">Velocity</span> Distribution Function (VDF) of He+ pickup ions shows two distinct populations that reflect the sources of these ions. The highly suprathermal population is the result of the ionization and pickup of almost resting interstellar neutrals that are injected into the solar wind as a highly anisotropic torus distribution. The nearly thermalized population is centered around the solar wind bulk speed and is mainly attributed to inner-source pickup ions that originate in the inner heliosphere. It is generally believed that the initial torus distribution of interstellar pickup ions is rapidly isotropized by resonant wave-particle interactions, but recent <span class="hlt">observations</span> by Drews et al. (2015) of a torus-like VDF strongly limit this isotropization. This in turn means that more <span class="hlt">observational</span> data is needed to further characterize the kinetic behavior of pickup ions. In this study we use data from the Charge-Time-Of-Flight sensor on-board SOHO. As this sensor offers unrivaled counting statistics for He+ together with a sufficient mass-per-charge resolution it is well-suited for investigating the He+ VDF on comparatively short timescales. We combine this data with the high resolution magnetic <span class="hlt">field</span> data from WIND via an extrapolation to the location of SOHO. With this combination of instruments we investigate the He+ VDF for time periods of different solar wind speeds, magnetic <span class="hlt">field</span> directions, and wave power. We find a systematic trend of the short-term He+ VDF with these parameters. Especially by varying the considered magnetic <span class="hlt">field</span> directions we <span class="hlt">observe</span> a 1D projection of the anisotropic torus-like VDF. In addition, we investigate stream interaction regions and coronal mass ejections. In the latter we <span class="hlt">observe</span> an excess of inner-source He+ that is accompanied by a significant increase of heavy pickup ion count rates. This may be linked to the as yet ill understood production mechanism of inner-source pickup ions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950034618&hterms=local+linear&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dlocal%2Blinear','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950034618&hterms=local+linear&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dlocal%2Blinear"><span>Comparison of large-scale structures and <span class="hlt">velocities</span> in the local universe</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yahil, Amos</p> <p>1994-01-01</p> <p>Comparison of the large-scale density and <span class="hlt">velocity</span> <span class="hlt">fields</span> in the local universe shows detailed agreement, strengthening the standard paradigm of the gravitational origin of these structures. Quantitative analysis can determine the cosmological density parameter, Omega, and biasing factor, b; there is virtually no sensitivity in any local analyses to the cosmologial constant, lambda. Comparison of the dipole anisotropy of the cosmic microwave background with the acceleration due to the Infrared Astronomy Satellite (IRAS) galaxies puts the linear growth factor in the range beta approximately equals Omega (exp 0.6)/b = 0.6(+0.7/-0.3) (95% confidence). A direct comparison of the density and <span class="hlt">velocity</span> <span class="hlt">fields</span> of nearby galaxies gives beta = 1.3 (+0.7/-0.6), and from nonlinear analysis the weaker limit (Omega greater than 0.45 for b greater than 0.5 (again 95% confidence). A tighter limit (Omega greater than 0.3 (4-6 sigma)), is obtained by a reconstruction of the probability distribution function of the initial fluctuations from which the structures <span class="hlt">observed</span> today arose. The last two methods depend critically on the smooth <span class="hlt">velocity</span> <span class="hlt">field</span> determined from the <span class="hlt">observed</span> <span class="hlt">velocities</span> of nearby galaxies by the POTENT method. A new analysis of these <span class="hlt">velocities</span>, with more than three times the data used to obtain the above quoted results, is now underway and promises to tighten the uncertainties considerably, as well as reduce systematic bias.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.S31A0787C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.S31A0787C"><span>Shear-wave <span class="hlt">velocities</span> beneath the Harrat Rahat volcanic <span class="hlt">field</span>, Saudi Arabia, using ambient seismic noise analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Civilini, F.; Mooney, W.; Savage, M. K.; Townend, J.; Zahran, H. M.</p> <p>2017-12-01</p> <p>We present seismic shear-<span class="hlt">velocities</span> for Harrat Rahat, a Cenozoic bimodal alkaline volcanic <span class="hlt">field</span> in west-central Saudi Arabia, using seismic tomography from natural ambient noise. This project is part of an overall effort by the Saudi Geological Survey and the United States Geological Survey to describe the subsurface structure and assess hazards within the Saudi Arabian shield. Volcanism at Harrat Rahat began approximately 10 Ma, with at least three pulses around 10, 5, and 2 Ma, and at least several pulses in the Quaternary from 1.9 Ma to the present. This area is instrumented by 14 broadband Nanometrics Trillium T120 instruments across an array aperture of approximately 130 kilometers. We used a year of recorded natural ambient noise to determine group and phase <span class="hlt">velocity</span> surface wave dispersion maps with a 0.1 decimal degree resolution for radial-radial, transverse-transverse, and vertical-vertical components of the empirical Green's function. A grid-search method was used to carry out 1D shear-<span class="hlt">velocity</span> inversions at each latitude-longitude point and the results were interpolated to produce pseudo-3D shear <span class="hlt">velocity</span> models. The dispersion maps resolved a zone of slow surface wave <span class="hlt">velocity</span> south-east of the city of Medina spatially correlated with the 1256 CE eruption. A crustal layer interface at approximately 20 km depth was determined by the inversions for all components, matching the results of prior seismic-refraction studies. Cross-sections of the 3D shear <span class="hlt">velocity</span> models were compared to gravity measurements obtained in the south-east edge of the <span class="hlt">field</span>. We found that measurements of low gravity qualitatively correlate with low values of shear-<span class="hlt">velocity</span> below 20 km along the cross-section profile. We apply these methods to obtain preliminary tomography results on the entire Arabian Shield.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20070034836','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20070034836"><span>Effect of Temperature on Jet <span class="hlt">Velocity</span> Spectra</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bridges, James E.; Wernet, Mark P.</p> <p>2007-01-01</p> <p>Statistical jet noise prediction codes that accurately predict spectral directivity for both cold and hot jets are highly sought both in industry and academia. Their formulation, whether based upon manipulations of the Navier-Stokes equations or upon heuristic arguments, require substantial experimental <span class="hlt">observation</span> of jet turbulence statistics. Unfortunately, the statistics of most interest involve the space-time correlation of flow quantities, especially <span class="hlt">velocity</span>. Until the last 10 years, all turbulence statistics were made with single-point probes, such as hotwires or laser Doppler anemometry. Particle image velocimetry (PIV) brought many new insights with its ability to measure <span class="hlt">velocity</span> <span class="hlt">fields</span> over large regions of jets simultaneously; however, it could not measure <span class="hlt">velocity</span> at rates higher than a few <span class="hlt">fields</span> per second, making it unsuitable for obtaining temporal spectra and correlations. The development of time-resolved PIV, herein called TR-PIV, has removed this limitation, enabling measurement of <span class="hlt">velocity</span> <span class="hlt">fields</span> at high resolution in both space and time. In this paper, ground-breaking results from the application of TR-PIV to single-flow hot jets are used to explore the impact of heat on turbulent statistics of interest to jet noise models. First, a brief summary of validation studies is reported, undertaken to show that the new technique produces the same trusted results as hotwire at cold, low-speed jets. Second, <span class="hlt">velocity</span> spectra from cold and hot jets are compared to see the effect of heat on the spectra. It is seen that heated jets possess 10 percent more turbulence intensity compared to the unheated jets with the same <span class="hlt">velocity</span>. The spectral shapes, when normalized using Strouhal scaling, are insensitive to temperature if the stream-wise location is normalized relative to the potential core length. Similarly, second order <span class="hlt">velocity</span> correlations, of interest in modeling of jet noise sources, are also insensitive to temperature as well.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AcAau.102...89M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AcAau.102...89M"><span>Magnetometer-only attitude and angular <span class="hlt">velocity</span> filtering estimation for attitude changing spacecraft</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ma, Hongliang; Xu, Shijie</p> <p>2014-09-01</p> <p>This paper presents an improved real-time sequential filter (IRTSF) for magnetometer-only attitude and angular <span class="hlt">velocity</span> estimation of spacecraft during its attitude changing (including fast and large angular attitude maneuver, rapidly spinning or uncontrolled tumble). In this new magnetometer-only attitude determination technique, both attitude dynamics equation and first time derivative of measured magnetic <span class="hlt">field</span> vector are directly leaded into filtering equations based on the traditional single vector attitude determination method of gyroless and real-time sequential filter (RTSF) of magnetometer-only attitude estimation. The process noise model of IRTSF includes attitude kinematics and dynamics equations, and its measurement model consists of magnetic <span class="hlt">field</span> vector and its first time derivative. The <span class="hlt">observability</span> of IRTSF for small or large angular <span class="hlt">velocity</span> changing spacecraft is evaluated by an improved Lie-Differentiation, and the degrees of <span class="hlt">observability</span> of IRTSF for different initial estimation errors are analyzed by the condition number and a solved covariance matrix. Numerical simulation results indicate that: (1) the attitude and angular <span class="hlt">velocity</span> of spacecraft can be estimated with sufficient accuracy using IRTSF from magnetometer-only data; (2) compared with that of RTSF, the estimation accuracies and <span class="hlt">observability</span> degrees of attitude and angular <span class="hlt">velocity</span> using IRTSF from magnetometer-only data are both improved; and (3) universality: the IRTSF of magnetometer-only attitude and angular <span class="hlt">velocity</span> estimation is <span class="hlt">observable</span> for any different initial state estimation error vector.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120015583','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120015583"><span>Measurements of Dendritic Growth <span class="hlt">Velocities</span> in Undercooled Melts of Pure Nickel Under Static Magnetic <span class="hlt">Fields</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Gao, Jianrong; Zhang, Zongning; Zhang, Yingjie</p> <p>2012-01-01</p> <p>Dendritic growth <span class="hlt">velocities</span> in undercooled melts of pure Ni have been intensively studied over the past fifty years. However, the literature data are at marked variance with the prediction of the widely accepted model for rapid dendritic growth both at small and at large undercoolings. In the present work, bulk melts of pure Ni samples of high purity were undercooled by glass fluxing treatment under a static magnetic <span class="hlt">field</span>. The recalescence processes of the samples at different undercoolings were recorded using a high-speed camera, and were modeled using a software to determine the dendritic growth <span class="hlt">velocities</span>. The present data confirmed the effect of melt flow on dendritic growth <span class="hlt">velocities</span> at undercoolings below 100 K. A comparison of the present data with previous measurements on a lower purity material suggested an effect of impurities on dendritic growth <span class="hlt">velocities</span> at undercoolings larger than 200 K as well.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=253162','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=253162"><span>A simple method to estimate threshold friction <span class="hlt">velocity</span> of wind erosion in the <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p>Nearly all wind erosion models require the specification of threshold friction <span class="hlt">velocity</span> (TFV). Yet determining TFV of wind erosion in <span class="hlt">field</span> conditions is difficult as it depends on both soil characteristics and distribution of vegetation or other roughness elements. While several reliable methods ha...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AdSpR..58..847W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AdSpR..58..847W"><span>The distribution of spectral index of magnetic <span class="hlt">field</span> and ion <span class="hlt">velocity</span> in Pi2 frequency band in BBFs: THEMIS statistics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Q.; Du, A. M.; Volwerk, M.; Wang, G. Q.</p> <p>2016-09-01</p> <p>A statistical study of the THEMIS FGM and ESA data is performed on turbulence of magnetic <span class="hlt">field</span> and <span class="hlt">velocity</span> for 218 selected 12 min intervals in BBFs. The spectral index α in the frequency range of 0.005-0.06 Hz are Gaussian distributions. The peaks indexes of total ion <span class="hlt">velocity</span> Vi and parallel <span class="hlt">velocity</span> V‖ are 1.95 and 2.07 nearly the spectral index of intermittent low frequency turbulence with large amplitude. However, most probable α of perpendicular <span class="hlt">velocity</span> V⊥ is about 1.75. It is a little bigger than 5/3 of Kolmogorov (1941). The peak indexes of total magnetic <span class="hlt">field</span> BT is 1.70 similar to V⊥. Compression magnetic <span class="hlt">field</span> B‖ are 1.85 which is smaller than 2 and bigger than 5/3 of Kolmogorov (1941). The most probable spectral index of shear B⊥ is about 1.44 which is close to 3/2 of Kraichnan (1965). Max V⊥ have little effect on the power magnitude of VT and V‖ but is positively correlated to spectral index of V⊥. The spectral power of BT, B‖ and B⊥ increase with max perpendicular <span class="hlt">velocity</span> but spectral indexes of them are negatively correlated to V⊥. The spectral index and the spectral power of magnetic <span class="hlt">field</span> over the frequency interval 0.005-0.06 Hz is very different from that over 0.08-1 Hz.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.475..379M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.475..379M"><span>Constraining the optical depth of galaxies and <span class="hlt">velocity</span> bias with cross-correlation between the kinetic Sunyaev-Zeldovich effect and the peculiar <span class="hlt">velocity</span> <span class="hlt">field</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, Yin-Zhe; Gong, Guo-Dong; Sui, Ning; He, Ping</p> <p>2018-03-01</p> <p>We calculate the cross-correlation function < (Δ T/T)({v}\\cdot \\hat{n}/σ _v) > between the kinetic Sunyaev-Zeldovich (kSZ) effect and the reconstructed peculiar <span class="hlt">velocity</span> <span class="hlt">field</span> using linear perturbation theory, with the aim of constraining the optical depth τ and peculiar <span class="hlt">velocity</span> bias of central galaxies with Planck data. We vary the optical depth τ and the <span class="hlt">velocity</span> bias function bv(k) = 1 + b(k/k0)n, and fit the model to the data, with and without varying the calibration parameter y0 that controls the vertical shift of the correlation function. By constructing a likelihood function and constraining the τ, b and n parameters, we find that the quadratic power-law model of <span class="hlt">velocity</span> bias, bv(k) = 1 + b(k/k0)2, provides the best fit to the data. The best-fit values are τ = (1.18 ± 0.24) × 10-4, b=-0.84^{+0.16}_{-0.20} and y0=(12.39^{+3.65}_{-3.66})× 10^{-9} (68 per cent confidence level). The probability of b > 0 is only 3.12 × 10-8 for the parameter b, which clearly suggests a detection of scale-dependent <span class="hlt">velocity</span> bias. The fitting results indicate that the large-scale (k ≤ 0.1 h Mpc-1) <span class="hlt">velocity</span> bias is unity, while on small scales the bias tends to become negative. The value of τ is consistent with the stellar mass-halo mass and optical depth relationship proposed in the literature, and the negative <span class="hlt">velocity</span> bias on small scales is consistent with the peak background split theory. Our method provides a direct tool for studying the gaseous and kinematic properties of galaxies.</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('http://adsabs.harvard.edu/abs/2016APS..DFDA31005N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DFDA31005N"><span>Application of photogrammetry to transforming PIV-acquired <span class="hlt">velocity</span> <span class="hlt">fields</span> to a moving-body coordinate system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nikoueeyan, Pourya; Naughton, Jonathan</p> <p>2016-11-01</p> <p>Particle Image Velocimetry is a common choice for qualitative and quantitative characterization of unsteady flows associated with moving bodies (e.g. pitching and plunging airfoils). Characterizing the separated flow behavior is of great importance in understanding the flow physics and developing predictive reduced-order models. In most studies, the model under investigation moves within a fixed camera <span class="hlt">field</span>-of-view, and vector <span class="hlt">fields</span> are calculated based on this fixed coordinate system. To better characterize the genesis and evolution of vortical structures in these unsteady flows, the <span class="hlt">velocity</span> <span class="hlt">fields</span> need to be transformed into the moving-body frame of reference. Data converted to this coordinate system allow for a more detailed analysis of the flow <span class="hlt">field</span> using advanced statistical tools. In this work, a pitching NACA0015 airfoil has been used to demonstrate the capability of photogrammetry for such an analysis. Photogrammetry has been used first to locate the airfoil within the image and then to determine an appropriate mask for processing the PIV data. The photogrammetry results are then further used to determine the rotation matrix that transforms the <span class="hlt">velocity</span> <span class="hlt">fields</span> to airfoil coordinates. Examples of the important capabilities such a process enables are discussed. P. Nikoueeyan is supported by a fellowship from the University of Wyoming's Engineering Initiative.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.G42A..03K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.G42A..03K"><span>A new GNSS <span class="hlt">velocity</span> <span class="hlt">field</span> for Fennoscandia and comparison to GIA models (Invited)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kierulf, H. P.; Simpson, M. J.; Steffen, H.; Lidberg, M.</p> <p>2013-12-01</p> <p>In Fennoscandia, the process of Glacial Isostatic Adjustment (GIA) causes ongoing crustal deformation. The vertical and horizontal movements of the Earth can be measured to a high degree of precision using Global Navigation Satellite System (GNSS). The GNSS network in Fennoscandia has gradually been established since the early 1990s and today contains a dense network well suited for geophysical studies and especially GIA. We will present new <span class="hlt">velocity</span> estimates for the Fennoscandian and North-European GNSS network using the processing package GAMIT/GLOBK. GNSS measurements have proved to be a good tool to constrain and validate GIA models. However, reference frame uncertainties, plate tectonics as well as intra-plate deformations might decontaminate the results. Different ITRFs have had large discrepancies, especially in the TZ-component, which have made the geophysical interpretation of GNSS results difficult. In GIA areas the uncertainties in the TZ component almost directly affect the height component which makes constraining of GIA models less reliable. Plate tectonics introduces large horizontal <span class="hlt">velocities</span> which are hard to distinguish from horizontal GIA-induced <span class="hlt">velocities</span>. We will present a new approach where our GNSS <span class="hlt">velocity</span> <span class="hlt">field</span> is directly realized in a GIA frame. With this approach, the effect of systematic errors in the reference frames and 'biasing' signal from the plate tectonics will be reduced to a minimum for our GIA results. Moreover, we are able to provide consistent GIA-free plate <span class="hlt">velocities</span> for the Eurasian plate.</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/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/2017A%26A...602A..60G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017A%26A...602A..60G"><span>Flare-induced changes of the photospheric magnetic <span class="hlt">field</span> in a δ-spot deduced from ground-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>Gömöry, P.; Balthasar, H.; Kuckein, C.; Koza, J.; Veronig, A. M.; González Manrique, S. J.; Kučera, A.; Schwartz, P.; Hanslmeier, A.</p> <p>2017-06-01</p> <p>Aims: Changes of the magnetic <span class="hlt">field</span> and the line-of-sight <span class="hlt">velocities</span> in the photosphere are being reported for an M-class flare that originated at a δ-spot belonging to active region NOAA 11865. Methods: High-resolution ground-based near-infrared spectropolarimetric <span class="hlt">observations</span> were acquired simultaneously in two photospheric spectral lines, Fe I 10783 Å and Si I 10786 Å, with the Tenerife Infrared Polarimeter at the Vacuum Tower Telescope (VTT) in Tenerife on 2013 October 15. The <span class="hlt">observations</span> covered several stages of the M-class flare. Inversions of the full-Stokes vector of both lines were carried out and the results were put into context using (extreme)-ultraviolet filtergrams from the Solar Dynamics Observatory (SDO). Results: The active region showed high flaring activity during the whole <span class="hlt">observing</span> period. After the M-class flare, the longitudinal magnetic <span class="hlt">field</span> did not show significant changes along the polarity inversion line (PIL). However, an enhancement of the transverse magnetic <span class="hlt">field</span> of approximately 550 G was found that bridges the PIL and connects umbrae of opposite polarities in the δ-spot. At the same time, a newly formed system of loops appeared co-spatially in the corona as seen in 171 Å filtergrams of the Atmospheric Imaging Assembly (AIA) on board SDO. However, we cannot exclude that the magnetic connection between the umbrae already existed in the upper atmosphere before the M-class flare and became visible only later when it was filled with hot plasma. The photospheric Doppler <span class="hlt">velocities</span> show a persistent upflow pattern along the PIL without significant changes due to the flare. Conclusions: The increase of the transverse component of the magnetic <span class="hlt">field</span> after the flare together with the newly formed loop system in the corona support recent predictions of flare models and flare <span class="hlt">observations</span>. The movie associated to Figs. 4 and 5 is available at http://www.aanda.org</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19760016074','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19760016074"><span>A submerged singularity method for calculating potential flow <span class="hlt">velocities</span> at arbitrary near-<span class="hlt">field</span> points</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Maskew, B.</p> <p>1976-01-01</p> <p>A discrete singularity method has been developed for calculating the potential flow around two-dimensional airfoils. The objective was to calculate <span class="hlt">velocities</span> at any arbitrary point in the flow <span class="hlt">field</span>, including points that approach the airfoil surface. That objective was achieved and is demonstrated here on a Joukowski airfoil. The method used combined vortices and sources ''submerged'' a small distance below the airfoil surface and incorporated a near-<span class="hlt">field</span> subvortex technique developed earlier. When a <span class="hlt">velocity</span> calculation point approached the airfoil surface, the number of discrete singularities effectively increased (but only locally) to keep the point just outside the error region of the submerged singularity discretization. The method could be extended to three dimensions, and should improve nonlinear methods, which calculate interference effects between multiple wings, and which include the effects of force-free trailing vortex sheets. The capability demonstrated here would extend the scope of such calculations to allow the close approach of wings and vortex sheets (or vortices).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5889303','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5889303"><span>Transformations Based on Continuous Piecewise-Affine <span class="hlt">Velocity</span> <span class="hlt">Fields</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>Freifeld, Oren; Hauberg, Søren; Batmanghelich, Kayhan; Fisher, Jonn W.</p> <p>2018-01-01</p> <p>We propose novel finite-dimensional spaces of well-behaved ℝn → ℝn transformations. The latter are obtained by (fast and highly-accurate) integration of continuous piecewise-affine <span class="hlt">velocity</span> <span class="hlt">fields</span>. The proposed method is simple yet highly expressive, effortlessly handles optional constraints (e.g., volume preservation and/or boundary conditions), and supports convenient modeling choices such as smoothing priors and coarse-to-fine analysis. Importantly, the proposed approach, partly due to its rapid likelihood evaluations and partly due to its other properties, facilitates tractable inference over rich transformation spaces, including using Markov-Chain Monte-Carlo methods. Its applications include, but are not limited to: monotonic regression (more generally, optimization over monotonic functions); modeling cumulative distribution functions or histograms; time-warping; image warping; image registration; real-time diffeomorphic image editing; data augmentation for image classifiers. Our GPU-based code is publicly available. PMID:28092517</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1367523-transformations-based-continuous-piecewise-affine-velocity-fields','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1367523-transformations-based-continuous-piecewise-affine-velocity-fields"><span>Transformations based on continuous piecewise-affine <span class="hlt">velocity</span> <span class="hlt">fields</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Freifeld, Oren; Hauberg, Soren; Batmanghelich, Kayhan; ...</p> <p>2017-01-11</p> <p>Here, we propose novel finite-dimensional spaces of well-behaved Rn → Rn transformations. The latter are obtained by (fast and highly-accurate) integration of continuous piecewise-affine <span class="hlt">velocity</span> <span class="hlt">fields</span>. The proposed method is simple yet highly expressive, effortlessly handles optional constraints (e.g., volume preservation and/or boundary conditions), and supports convenient modeling choices such as smoothing priors and coarse-to-fine analysis. Importantly, the proposed approach, partly due to its rapid likelihood evaluations and partly due to its other properties, facilitates tractable inference over rich transformation spaces, including using Markov-Chain Monte-Carlo methods. Its applications include, but are not limited to: monotonic regression (more generally, optimization overmore » monotonic functions); modeling cumulative distribution functions or histograms; time-warping; image warping; image registration; real-time diffeomorphic image editing; data augmentation for image classifiers. Our GPU-based code is publicly available.« less</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> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22251044-remarks-regularity-criteria-three-dimensional-magnetohydrodynamics-system-terms-two-velocity-field-components','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22251044-remarks-regularity-criteria-three-dimensional-magnetohydrodynamics-system-terms-two-velocity-field-components"><span>Remarks on the regularity criteria of three-dimensional magnetohydrodynamics system in terms of two <span class="hlt">velocity</span> <span class="hlt">field</span> components</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>Yamazaki, Kazuo</p> <p>2014-03-15</p> <p>We study the three-dimensional magnetohydrodynamics system and obtain its regularity criteria in terms of only two <span class="hlt">velocity</span> vector <span class="hlt">field</span> components eliminating the condition on the third component completely. The proof consists of a new decomposition of the four nonlinear terms of the system and estimating a component of the magnetic vector <span class="hlt">field</span> in terms of the same component of the <span class="hlt">velocity</span> vector <span class="hlt">field</span>. This result may be seen as a component reduction result of many previous works [C. He and Z. Xin, “On the regularity of weak solutions to the magnetohydrodynamic equations,” J. Differ. Equ. 213(2), 234–254 (2005); Y. Zhou,more » “Remarks on regularities for the 3D MHD equations,” Discrete Contin. Dyn. Syst. 12(5), 881–886 (2005)].« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.H51S..07M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.H51S..07M"><span>Near-<span class="hlt">field</span> Oblique Remote Sensing of Stream Water-surface Elevation, Slope, and Surface <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>Minear, J. T.; Kinzel, P. J.; Nelson, J. M.; McDonald, R.; Wright, S. A.</p> <p>2014-12-01</p> <p>A major challenge for estimating discharges during flood events or in steep channels is the difficulty and hazard inherent in obtaining in-stream measurements. One possible solution is to use near-<span class="hlt">field</span> remote sensing to obtain simultaneous water-surface elevations, slope, and surface <span class="hlt">velocities</span>. In this test case, we utilized Terrestrial Laser Scanning (TLS) to remotely measure water-surface elevations and slope in combination with surface <span class="hlt">velocities</span> estimated from particle image velocimetry (PIV) obtained by video-camera and/or infrared camera. We tested this method at several sites in New Mexico and Colorado using independent validation data consisting of in-channel measurements from survey-grade GPS and Acoustic Doppler Current Profiler (ADCP) instruments. Preliminary results indicate that for relatively turbid or steep streams, TLS collects tens of thousands of water-surface elevations and slopes in minutes, much faster than conventional means and at relatively high precision, at least as good as continuous survey-grade GPS measurements. Estimated surface <span class="hlt">velocities</span> from this technique are within 15% of measured <span class="hlt">velocity</span> magnitudes and within 10 degrees from the measured <span class="hlt">velocity</span> direction (using extrapolation from the shallowest bin of the ADCP measurements). Accurately aligning the PIV results into Cartesian coordinates appears to be one of the main sources of error, primarily due to the sensitivity at these shallow oblique look angles and the low numbers of stationary objects for rectification. Combining remotely-sensed water-surface elevations, slope, and surface <span class="hlt">velocities</span> produces simultaneous <span class="hlt">velocity</span> measurements from a large number of locations in the channel and is more spatially extensive than traditional <span class="hlt">velocity</span> measurements. These factors make this technique useful for improving estimates of flow measurements during flood flows and in steep channels while also decreasing the difficulty and hazard associated with making measurements in these</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017CoGG...47..261S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017CoGG...47..261S"><span>Magnetic and <span class="hlt">velocity</span> <span class="hlt">fields</span> in a dynamo operating at extremely small Ekman and magnetic Prandtl numbers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Šimkanin, Ján; Kyselica, Juraj</p> <p>2017-12-01</p> <p>Numerical simulations of the geodynamo are becoming more realistic because of advances in computer technology. Here, the geodynamo model is investigated numerically at the extremely low Ekman and magnetic Prandtl numbers using the PARODY dynamo code. These parameters are more realistic than those used in previous numerical studies of the geodynamo. Our model is based on the Boussinesq approximation and the temperature gradient between upper and lower boundaries is a source of convection. This study attempts to answer the question how realistic the geodynamo models are. Numerical results show that our dynamo belongs to the strong-<span class="hlt">field</span> dynamos. The generated magnetic <span class="hlt">field</span> is dipolar and large-scale while convection is small-scale and sheet-like flows (plumes) are preferred to a columnar convection. Scales of magnetic and <span class="hlt">velocity</span> <span class="hlt">fields</span> are separated, which enables hydromagnetic dynamos to maintain the magnetic <span class="hlt">field</span> at the low magnetic Prandtl numbers. The inner core rotation rate is lower than that in previous geodynamo models. On the other hand, dimensional magnitudes of <span class="hlt">velocity</span> and magnetic <span class="hlt">fields</span> and those of the magnetic and viscous dissipation are larger than those expected in the Earth's core due to our parameter range chosen.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1815841T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1815841T"><span>Systematic Variability of the He+ Pickup Ion <span class="hlt">Velocity</span> Distribution Function <span class="hlt">Observed</span> with SOHO/CELIAS/CTOF</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Taut, Andreas; Drews, Christian; Berger, Lars; Wimmer-Schweingruber, Robert</p> <p>2016-04-01</p> <p>The 1D <span class="hlt">Velocity</span> Distribution Function (VDF) of He+ pickup ions shows two distinct populations that reflect the sources of these ions. The highly suprathermal population is the result of the ionization and pickup of almost resting interstellar neutrals that are injected into the solar wind as a highly anisotropic torus distribution. The nearly thermalized population is centered around the solar wind bulk speed and is mainly attributed to inner-source pickup ions that originate in the inner heliosphere. Current pickup ion models assume a rapid isotropization of the initial VDF by resonant wave-particle interactions, but recent <span class="hlt">observations</span> by Drews et al. (2015) of a torus-like VDF strongly limit this isotropization. This in turn means that more <span class="hlt">observational</span> data is needed to further characterize the kinetic behavior of pickup ions. The Charge-Time-Of-Flight sensor on-board SOHO offers unrivaled counting statistics for He+ and a sufficient mass-per-charge resolution. Thus, the He+ VDF can be <span class="hlt">observed</span> on comparatively short timescales. We combine this data with the magnetic <span class="hlt">field</span> data from WIND via an extrapolation to the location of SOHO. On the one hand we investigate the 1D VDF of He+ pickup ions with respect to different magnetic <span class="hlt">field</span> orientations. Our findings complement on previous studies with other instruments that show an anisotropy of the VDF that is linked to the initial torus VDF. On the other hand we find a significant modification of the VDF during stream-interaction region. This may be linked to a different cooling behaviour in these regions and/or the absence of inner-source He+ during these times. Here, we report on our preliminary results.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhPl...24h2308C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhPl...24h2308C"><span>Magnetic and <span class="hlt">velocity</span> fluctuations from nonlinearly coupled tearing modes in the reversed <span class="hlt">field</span> pinch with and without the reversal surface</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Craig, D.; Martin, D.; Den Hartog, D. J.; Nornberg, M. D.; Reusch, J. A.</p> <p>2017-08-01</p> <p>We investigate the role of poloidal mode number m = 0 fluctuations on m = 1 <span class="hlt">velocity</span> and magnetic <span class="hlt">field</span> fluctuations in the Reversed <span class="hlt">Field</span> Pinch (RFP). Removing the m = 0 resonant surface in the Madison Symmetric Torus (MST), results in suppressed m = 0 activity without a reduction in m = 1 magnetic activity. However, the m = 1 <span class="hlt">velocity</span> fluctuations and fluctuation-induced mean emf are reduced as m = 0 modes are suppressed. <span class="hlt">Velocity</span> fluctuations are measured directly using fast Doppler spectroscopy. Similar results are seen in visco-resistive MHD simulation with the DEBS code. An artificial line-averaged <span class="hlt">velocity</span> diagnostic is developed for DEBS simulations to facilitate direct comparisons with experimental measurements. The sensitivity of the m = 1 <span class="hlt">velocity</span> fluctuations and corresponding emf to changes in m = 0 mode activity is a feature of tearing modes in the nonlinear regime with a spectrum of interacting modes. These results have implications for RFP sustainment strategies and inform our understanding of the role of magnetic turbulence in astrophysical contexts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SoPh..293...25N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SoPh..293...25N"><span>Understanding the Internal Magnetic <span class="hlt">Field</span> Configurations of ICMEs Using More than 20 Years of Wind <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>Nieves-Chinchilla, T.; Vourlidas, A.; Raymond, J. C.; Linton, M. G.; Al-haddad, N.; Savani, N. P.; Szabo, A.; Hidalgo, M. A.</p> <p>2018-02-01</p> <p>The magnetic topology, structure, and geometry of the magnetic obstacles embedded within interplanetary coronal mass ejections (ICMEs) are not yet fully and consistently described by in situ models and reconstruction techniques. The main goal of this work is to better understand the status of the internal magnetic <span class="hlt">field</span> of ICMEs and to explore in situ signatures to identify clues to develop a more accurate and reliable in situ analytical models. We take advantage of more than 20 years of Wind <span class="hlt">observations</span> of transients at 1 AU to compile a comprehensive database of ICMEs through three solar cycles, from 1995 to 2015. The catalog is publicly available at wind.gsfc.nasa.gov and is fully described in this article. We identify and collect the properties of 337 ICMEs, of which 298 show organized magnetic <span class="hlt">field</span> signatures. To allow for departures from idealized magnetic configurations, we introduce the term "magnetic obstacle" (MO) to signify the possibility of more complex configurations. To quantify the asymmetry of the magnetic <span class="hlt">field</span> strength profile within these events, we introduce the distortion parameter (DiP) and calculate the expansion <span class="hlt">velocity</span> within the magnetic obstacle. Circular-cylindrical geometry is assumed when the magnetic <span class="hlt">field</span> strength displays a symmetric profile. We perform a statistical study of these two parameters and find that only 35% of the events show symmetric magnetic profiles and a low enough expansion <span class="hlt">velocity</span> to be compatible with the assumption of an idealized cylindrical static flux rope, and that 41% of the events do not show the expected relationship between expansion and magnetic <span class="hlt">field</span> compression in the front, with the maximum magnetic <span class="hlt">field</span> closer to the first encounter of the spacecraft with the magnetic obstacle; 18% show contractions ( i.e. apparent negative expansion <span class="hlt">velocity</span>), and 30% show magnetic <span class="hlt">field</span> compression in the back. We derive an empirical relation between DiP and expansion <span class="hlt">velocity</span> that is the first step toward</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017WRR....53.7521J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017WRR....53.7521J"><span>Remote determination of the <span class="hlt">velocity</span> index and mean streamwise <span class="hlt">velocity</span> profiles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Johnson, E. D.; Cowen, E. A.</p> <p>2017-09-01</p> <p>When determining volumetric discharge from surface measurements of currents in a river or open channel, the <span class="hlt">velocity</span> index is typically used to convert surface <span class="hlt">velocities</span> to depth-averaged <span class="hlt">velocities</span>. The <span class="hlt">velocity</span> index is given by, k=Ub/Usurf, where Ub is the depth-averaged <span class="hlt">velocity</span> and Usurf is the local surface <span class="hlt">velocity</span>. The USGS (United States Geological Survey) standard value for this coefficient, k = 0.85, was determined from a series of laboratory experiments and has been widely used in the <span class="hlt">field</span> and in laboratory measurements of volumetric discharge despite evidence that the <span class="hlt">velocity</span> index is site-specific. Numerous studies have documented that the <span class="hlt">velocity</span> index varies with Reynolds number, flow depth, and relative bed roughness and with the presence of secondary flows. A remote method of determining depth-averaged <span class="hlt">velocity</span> and hence the <span class="hlt">velocity</span> index is developed here. The technique leverages the findings of Johnson and Cowen (2017) and permits remote determination of the <span class="hlt">velocity</span> power-law exponent thereby, enabling remote prediction of the vertical structure of the mean streamwise <span class="hlt">velocity</span>, the depth-averaged <span class="hlt">velocity</span>, and the <span class="hlt">velocity</span> index.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2000ECSS...50..185V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2000ECSS...50..185V"><span>Dynamics of Plant Flow Interactions for the Seagrass Amphibolis antarctica: <span class="hlt">Field</span> <span class="hlt">Observations</span> and 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>Verduin, J. J.; Backhaus, J. O.</p> <p>2000-02-01</p> <p>Seagrass canopies influence water flow partly as a consequence of their morphology. Amphibolis antarctica (Labill.) Sonder et Aschers. ex Aschers, an Australian endemic, is different morphologically from more-commonly studied blade-like seagrasses such as Zostera and Thalassia. <span class="hlt">Field</span> measurements and model predictions were used to characterize water flow within and above an A. antarctica meadow. A series of high resolution three-dimensional <span class="hlt">velocity</span> measurements were obtained within, above and adjacent to A. antarctica meadows at different heights above the seabed. <span class="hlt">Field</span> <span class="hlt">observations</span> on the effect of seagrass canopy on flow show an overall damping effect. Power spectra of the <span class="hlt">velocity</span> data revealed a reduction in energy from 500 (cm s -1) 2s -1to 10 (cm s -1) 2s -1within the canopy. Profiles of kinetic energy were calculated from in situ <span class="hlt">velocity</span> measurements at 5 cm increments from 10 cm to 80 cm above the seabed, within and above the seagrass canopy. There was an intensification of flow where the canopy structure was densest (approximately 40 cm above the seabed) and slightly above it. The baffling effect of the canopy was most effective 25 cm above the seabed: here the flow was reduced from 50 cm s -1at free surface to 2-5 cm s -1. A slight increase in flow within the canopy was seen 10 cm above the sediment due to reduced friction exerted by the lower leafless stems of the plants. A high resolution three-dimensional hydrodynamic model was coupled to a ten-layer canopy model for shallow coastal site dimensions. By applying different friction factors to various parts of the plant, mimicking its architecture, water flow was shown to be altered by the plant canopy according to its morphology. The derived computational results were in good agreement with the <span class="hlt">observed</span> in situ <span class="hlt">velocity</span> and kinetic energy changes. As a result of this study it is now possible to accurately predict plant-flow interactions determining pollen and particles distribution and dispersal.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.471.1181B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.471.1181B"><span>Characterization of the <span class="hlt">velocity</span> anisotropy of accreted globular clusters</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bianchini, P.; Sills, A.; Miholics, M.</p> <p>2017-10-01</p> <p>Galactic globular clusters (GCs) are believed to have formed in situ in the Galaxy as well as in dwarf galaxies later accreted on to the Milky Way. However, to date, there is no unambiguous signature to distinguish accreted GCs. Using specifically designed N-body simulations of GCs evolving in a variety of time-dependent tidal <span class="hlt">fields</span> (describing the potential of a dwarf galaxy-Milky Way merger), we analyse the effects imprinted on the internal kinematics of an accreted GC. In particular, we look at the evolution of the <span class="hlt">velocity</span> anisotropy. Our simulations show that at early phases, the <span class="hlt">velocity</span> anisotropy is determined by the tidal <span class="hlt">field</span> of the dwarf galaxy and subsequently the clusters will adapt to the new tidal environment, losing any signature of their original environment in a few relaxation times. At 10 Gyr, GCs exhibit a variety of <span class="hlt">velocity</span> anisotropy profiles, namely, isotropic <span class="hlt">velocity</span> distribution in the inner regions and either isotropy or radial/tangential anisotropy in the intermediate and outer regions. Independent of an accreted origin, the <span class="hlt">velocity</span> anisotropy primarily depends on the strength of the tidal <span class="hlt">field</span> cumulatively experienced by a cluster. Tangentially anisotropic clusters correspond to systems that have experienced stronger tidal <span class="hlt">fields</span> and are characterized by higher tidal filling factor, r50/rj ≳ 0.17, higher mass-loss ≳ 60 per cent and relaxation times trel ≲ 109 Gyr. Interestingly, we demonstrate that the presence of tidal tails can significantly contaminate the measurements of <span class="hlt">velocity</span> anisotropy when a cluster is <span class="hlt">observed</span> in projection. Our characterization of the <span class="hlt">velocity</span> anisotropy profiles in different tidal environments provides a theoretical benchmark for the interpretation of the unprecedented amount of three-dimensional kinematic data progressively available for Galactic GCs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940017134','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940017134"><span>Instantaneous <span class="hlt">velocity</span> <span class="hlt">field</span> imaging instrument for supersonic reacting flows</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Allen, M. G.; Davis, S. J.; Kessler, W. J.; Legner, H. H.; Mcmanus, K. R.; Mulhall, P. A.; Parker, T. E.; Sonnenfroh, D. M.</p> <p>1993-01-01</p> <p>The technical tasks conducted to develop and demonstrate a new gas <span class="hlt">velocity</span> measurement technique for high enthalpy reacting flows is described. The technique is based on Doppler-shifted Planar Laser-induced Fluorescence (PLIF) imaging of the OH radical. The imaging approach permits, in principle, single-shot measurements of the 2-D distribution of a single <span class="hlt">velocity</span> component in the measurement plane, and is thus a technique of choice for applications in high enthalpy transient flow facilities. In contrast to previous work in this area, the present program demonstrated an approach which modified the diagnostic technique to function under the constraints of practical flow conditions of engineering interest, rather than vice-versa. In order to accomplish the experimental demonstrations, the state-of-the-art in PLIF diagnostic techniques was advanced in several ways. Each of these tasks is described in detail and is intended to serve as a reference in supporting the transition of this new capability to the <span class="hlt">fielded</span> PLIF instruments now installed at several national test facilities. Among the new results of general interest in LlF-based flow diagnostics, a detailed set of the first measurements of the collisional broadening and shifting behavior of OH (1,0) band transitions in H7-air combustion environments is included. Such measurements are critical in the design of a successful strategy for PLIF <span class="hlt">velocity</span> imaging; they also relate to accurate concentration and temperature measurements, particularly in compressible flow regimes. Furthermore, the results shed new light on the fundamental relationship between broadening and energy transfer collisions in OH A(sup 2)Sigma(+)v(sup ') = 1. The first single-pulse, spectrally-resolved measurements of the output of common pulsed dye lasers were also produced during the course of this effort. As with the OH broadening measurements, these data are a significant aspect of a successful <span class="hlt">velocity</span> imaging strategy, and also have</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('https://ntrs.nasa.gov/search.jsp?R=19820058743&hterms=Big+Bang&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DBig%2BBang','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19820058743&hterms=Big+Bang&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DBig%2BBang"><span>H0, q0 and the local <span class="hlt">velocity</span> <span class="hlt">field</span>. [Hubble and deceleration constants in Big Bang expansion</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sandage, A.; Tammann, G. A.</p> <p>1982-01-01</p> <p>An attempt is made to find a systematic deviation from linearity for distances that are under the control of the Virgo cluster, and to determine the value of the mean random motion about the systematic flow, in order to improve the measurement of the Hubble and the deceleration constants. The <span class="hlt">velocity</span>-distance relation for large and intermediate distances is studied, and type I supernovae are calibrated relatively as distance indicators and absolutely to obtain a new value for the Hubble constant. Methods of determining the deceleration constant are assessed, including determination from direct measurement, mean luminosity density, virgocentric motion, and the time scale test. The very local <span class="hlt">velocity</span> <span class="hlt">field</span> is investigated, and a solution is preferred with a random peculiar radial <span class="hlt">velocity</span> of very nearby <span class="hlt">field</span> galaxies of 90-100 km/s, and a Virgocentric motion of the local group of 220 km/s, leading to an underlying expansion rate of 55, in satisfactory agreement with the global value.</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> </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/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://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('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/2016EGUGA..1814162C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1814162C"><span>On retrodictions of global mantle flow with assimilated surface <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>Colli, Lorenzo; Bunge, Hans-Peter; Schuberth, Bernhard S. A.</p> <p>2016-04-01</p> <p>Modeling past states of Earth's mantle and relating them to geologic <span class="hlt">observations</span> such as continental-scale uplift and subsidence is an effective method for testing mantle convection models. However, mantle convection is chaotic and two identical mantle models initialized with slightly different temperature <span class="hlt">fields</span> diverge exponentially in time until they become uncorrelated, thus limiting retrodictions (i.e., reconstructions of past states of Earth's mantle obtained using present information) to the recent past. We show with 3-D spherical mantle convection models that retrodictions of mantle flow can be extended significantly if knowledge of the surface <span class="hlt">velocity</span> <span class="hlt">field</span> is available. Assimilating surface <span class="hlt">velocities</span> produces in some cases negative Lyapunov times (i.e., e-folding times), implying that even a severely perturbed initial condition may evolve toward the reference state. A history of the surface <span class="hlt">velocity</span> <span class="hlt">field</span> for Earth can be obtained from past plate motion reconstructions for time periods of a mantle overturn, suggesting that mantle flow can be reconstructed over comparable times.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015GeoRL..42.8341C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015GeoRL..42.8341C"><span>On retrodictions of global mantle flow with assimilated surface <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>Colli, Lorenzo; Bunge, Hans-Peter; Schuberth, Bernhard S. A.</p> <p>2015-10-01</p> <p>Modeling past states of Earth's mantle and relating them to geologic <span class="hlt">observations</span> such as continental-scale uplift and subsidence is an effective method for testing mantle convection models. However, mantle convection is chaotic and two identical mantle models initialized with slightly different temperature <span class="hlt">fields</span> diverge exponentially in time until they become uncorrelated, thus limiting retrodictions (i.e., reconstructions of past states of Earth's mantle obtained using present information) to the recent past. We show with 3-D spherical mantle convection models that retrodictions of mantle flow can be extended significantly if knowledge of the surface <span class="hlt">velocity</span> <span class="hlt">field</span> is available. Assimilating surface <span class="hlt">velocities</span> produces in some cases negative Lyapunov times (i.e., e-folding times), implying that even a severely perturbed initial condition may evolve toward the reference state. A history of the surface <span class="hlt">velocity</span> <span class="hlt">field</span> for Earth can be obtained from past plate motion reconstructions for time periods of a mantle overturn, suggesting that mantle flow can be reconstructed over comparable times.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSM33E..08H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSM33E..08H"><span>Partition of Heating During Magnetic Reconnection: Role of Exhaust <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>Haggerty, C. C.; Shay, M.; Drake, J. F.; Phan, T.; Chasapis, A.; Cassak, P.; Malakit, K.</p> <p>2017-12-01</p> <p>The partition of released magnetic energy into ion and electron bulk flow and thermal energy is an important problem that has recently become under intense scrutiny in the magnetosphere and heliosphere. In the strong magnetic shear limit of magnetic reconnection (low guide <span class="hlt">field</span>), the production of counter-streaming beams due to magnetic <span class="hlt">field</span> line contraction plays an important role in heating the plasma. The contraction <span class="hlt">velocity</span> or outflow <span class="hlt">velocity</span> controls the magnitude of the heating. Although it is known that often the outflow <span class="hlt">velocity</span> is less than the upstream Alfvén speed, an understanding of why this is so is lacking. We show that the outflow <span class="hlt">velocity</span> in reconnection is reduced by the parallel ion exhaust temperature and derive a scaling relationship for this effect. This prediction is found to be consistent with both kinetic PIC simulations and MMS satellite <span class="hlt">observations</span>. This scaling for the outflow is then applied to a general theory for plasma heating during magnetic reconnection.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AAS...23115804R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AAS...23115804R"><span>WFIRST: Microlensing Parallax <span class="hlt">Observations</span> from K2 in the Exoplanet Microlensing <span class="hlt">Field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ranc, Clement; Radek Poleski, David Bennett, K2C9 Microlensing Science Experiment Team</p> <p>2018-01-01</p> <p>The recent explosion in our understanding of exoplanetary systems has been driven primarily by the Kepler mission, which has replaced radial <span class="hlt">velocities</span> as our main planet discovery method. While Kepler has provided a large sample of planets that will allow a robust statistical determination of the properties of exoplanets in close orbits about their host stars, the Kepler mission was stopped shortly after the start of its 5th year. This led to the Kepler 2 (K2) mission, which could <span class="hlt">observe</span> up to 18 different <span class="hlt">fields</span> in the ecliptic plane, including a fraction of the WFIRST microlensing <span class="hlt">field</span>. The K2 mission has focused on lower mass host stars and spending one <span class="hlt">observing</span> campaign in the Galactic bulge to make use of Kepler’s orbit to determine the masses and distances to microlensing systems via the microlensing parallax effect. These K2 Campaign 9 <span class="hlt">observations</span> help to develop the microlensing planet detection method, which will be employed by the WFIRST mission that will extend the statistical census of exoplanets to include low-mass planets in wide orbits. While the photometric light curve of a microlensing event <span class="hlt">observed</span> from the ground provides important constraints on the lens physical parameters, in many cases the lens mass and distance from Earth remain degenerated. The poster will show how simultaneous space- and ground-based <span class="hlt">observations</span> can break this mass-distance degeneracy. This method will be used for a fraction of the events <span class="hlt">observed</span> by WFIRST. Finally, the poster will present a new method to correct the K2 photometry from the correlated systematic noise. This investigation helps in characterizing the properties of the lens stars and source stars in one WFIRST <span class="hlt">field</span> with high extinction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018IJGMM..1550047G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018IJGMM..1550047G"><span>A tale of two <span class="hlt">velocities</span>: Threading versus slicing</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gharechahi, Razieh; Nouri-Zonoz, Mohammad; Tavanfar, Alireza</p> <p></p> <p>One of the important quantities in cosmology and astrophysics is the 3-<span class="hlt">velocity</span> of an object. Specifically, when the gravitational <span class="hlt">fields</span> are strong, one should require the employment of general relativity both in its definition and measurement. Looking into the literature for GR-based definitions of 3-<span class="hlt">velocity</span>, one usually finds different ad hoc definitions applied according to the case under consideration. Here, we introduce and analyze systematically the two principal definitions of 3-<span class="hlt">velocity</span> assigned to a test particle following the timelike trajectories in stationary spacetimes. These definitions are based on the 1 + 3 (threading) and 3 + 1 (slicing) spacetime decomposition formalisms and defined relative to two different sets of <span class="hlt">observers</span>. After showing that Synge’s definition of spatial distance and 3-<span class="hlt">velocity</span> is equivalent to those defined in the 1 + 3 (threading) formalism, we exemplify the differences between these two definitions by calculating them for particles in circular orbits in axially symmetric stationary spacetimes. Illustrating its geometric nature, the relative linear <span class="hlt">velocity</span> between the corresponding <span class="hlt">observers</span> is obtained in terms of the spacetime metric components. Circular particle orbits in the Kerr spacetime, as the prototype and the most well known of stationary spacetimes, are examined with respect to these definitions to highlight their <span class="hlt">observer</span>-dependent nature. We also examine the Kerr-NUT spacetime in which the NUT parameter, contributing to the off-diagonal terms in the metric, is mainly interpreted not as a rotation parameter but as a gravitomagnetic monopole charge. Finally, in a specific astrophysical setup which includes rotating black holes, it is shown how the local <span class="hlt">velocity</span> of an orbiting star could be related to its spectral line shifts measured by distant <span class="hlt">observers</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1366949','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1366949"><span><span class="hlt">Field</span>-gradient partitioning for fracture and frictional contact in the material point method: <span class="hlt">Field</span>-gradient partitioning for fracture and frictional contact in the material point method [Fracture and frictional contact in material point method using damage-<span class="hlt">field</span> gradients for <span class="hlt">velocity-field</span> partitioning</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>Homel, Michael A.; Herbold, Eric B.</p> <p></p> <p>Contact and fracture in the material point method require grid-scale enrichment or partitioning of material into distinct <span class="hlt">velocity</span> <span class="hlt">fields</span> to allow for displacement or <span class="hlt">velocity</span> discontinuities at a material interface. We present a new method which a kernel-based damage <span class="hlt">field</span> is constructed from the particle data. The gradient of this <span class="hlt">field</span> 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 <span class="hlt">velocity</span> <span class="hlt">field</span>, 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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1366949-field-gradient-partitioning-fracture-frictional-contact-material-point-method-field-gradient-partitioning-fracture-frictional-contact-material-point-method-fracture-frictional-contact-material-point-method-using-damage-field-gradients-velocity-field-partitioning','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1366949-field-gradient-partitioning-fracture-frictional-contact-material-point-method-field-gradient-partitioning-fracture-frictional-contact-material-point-method-fracture-frictional-contact-material-point-method-using-damage-field-gradients-velocity-field-partitioning"><span><span class="hlt">Field</span>-gradient partitioning for fracture and frictional contact in the material point method: <span class="hlt">Field</span>-gradient partitioning for fracture and frictional contact in the material point method [Fracture and frictional contact in material point method using damage-<span class="hlt">field</span> gradients for <span class="hlt">velocity-field</span> partitioning</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Homel, Michael A.; Herbold, Eric B.</p> <p>2016-08-15</p> <p>Contact and fracture in the material point method require grid-scale enrichment or partitioning of material into distinct <span class="hlt">velocity</span> <span class="hlt">fields</span> to allow for displacement or <span class="hlt">velocity</span> discontinuities at a material interface. We present a new method which a kernel-based damage <span class="hlt">field</span> is constructed from the particle data. The gradient of this <span class="hlt">field</span> 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 <span class="hlt">velocity</span> <span class="hlt">field</span>, 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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeoJI.213.1599L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeoJI.213.1599L"><span>Rayleigh wave group <span class="hlt">velocity</span> and shear wave <span class="hlt">velocity</span> structure in the San Francisco Bay region from ambient noise tomography</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Peng; Thurber, Clifford</p> <p>2018-06-01</p> <p>We derive new Rayleigh wave group <span class="hlt">velocity</span> models and a 3-D shear wave <span class="hlt">velocity</span> model of the upper crust in the San Francisco Bay region using an adaptive grid ambient noise tomography algorithm and 6 months of continuous seismic data from 174 seismic stations from multiple networks. The resolution of the group <span class="hlt">velocity</span> models is 0.1°-0.2° for short periods (˜3 s) and 0.3°-0.4° for long periods (˜10 s). The new shear wave <span class="hlt">velocity</span> model of the upper crust reveals a number of important structures. We find distinct <span class="hlt">velocity</span> contrasts at the Golden Gate segment of the San Andreas Fault, the West Napa Fault, central part of the Hayward Fault and southern part of the Calaveras Fault. Low shear wave <span class="hlt">velocities</span> are mainly located in Tertiary and Quaternary basins, for instance, La Honda Basin, Livermore Valley and the western and eastern edges of Santa Clara Valley. Low shear wave <span class="hlt">velocities</span> are also <span class="hlt">observed</span> at the Sonoma volcanic <span class="hlt">field</span>. Areas of high shear wave <span class="hlt">velocity</span> include the Santa Lucia Range, the Gabilan Range and Ben Lomond Plutons, and the Diablo Range, where Franciscan Complex or Silinian rocks are exposed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFMOS23B1185C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMOS23B1185C"><span>In-situ <span class="hlt">Observations</span> of Swash-zone Flow <span class="hlt">Velocities</span> and Sediment Transport on a Steep Beach</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chardon-Maldonado, P.; Puleo, J. A.; Figlus, J.</p> <p>2014-12-01</p> <p>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 <span class="hlt">velocity</span> 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 <span class="hlt">velocities</span> (u,v,w) were measured at each cross-shore location using a Nortek Vectrino profiling velocimeter that measured a 30 mm <span class="hlt">velocity</span> profile at 1 mm vertical increments at 100 Hz. These <span class="hlt">velocity</span> 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 <span class="hlt">velocity</span> 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 <span class="hlt">velocities</span> and bed shear stress estimates will be related to the <span class="hlt">observed</span> morphological change.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19930017891','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19930017891"><span>Determination of plate wave <span class="hlt">velocities</span> and diffuse <span class="hlt">field</span> decay rates with braod-band acousto-ultrasonic signals</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kautz, Harold E.</p> <p>1993-01-01</p> <p>Lowest symmetric and lowest antisymmetric plate wave modes were excited and identified in broad-band acousto-ultrasonic (AU) signals collected from various high temperature composite materials. Group <span class="hlt">velocities</span> have been determined for these nearly nondispersive modes. An algorithm has been developed and applied to determine phase <span class="hlt">velocities</span> and hence dispersion curves for the frequency ranges of the broad-band pulses. It is demonstrated that these data are sensitive to changes in the various stiffness moduli of the materials, in agreement by analogy, with the theoretical and experimental results of Tang and Henneke on fiber reinforced polymers. Diffuse <span class="hlt">field</span> decay rates have been determined in the same specimen geometries and AU configuration as for the plate wave measurements. These decay rates are of value in assessing degradation such as matrix cracking in ceramic matrix composites. In addition, we verify that diffuse <span class="hlt">field</span> decay rates respond to fiber/matrix interfacial shear strength and density in ceramic matrix composites. This work shows that <span class="hlt">velocity</span>/stiffness and decay rate measurements can be obtained in the same set of AU experiments for characterizing materials and in specimens with geometries useful for mechanical measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20000092053','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20000092053"><span>Photometric <span class="hlt">Observations</span> of 6000 Stars in the Cygnus <span class="hlt">Field</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Borucki, W.; Caldwell, D.; Koch, D.; Jenkins, J.; Ninkov, Z.</p> <p>1999-01-01</p> <p>A small photometer to detect transits by extrasolar planets has been assembled and is being tested at Lick Observatory on Mt. Hamilton, California. The Vulcan photometer is constructed from a 30 cm focal length, F/2.5 AeroEktar reconnaissance lens and Photometrics PXL16800 CCD camera. A spectral filter is used to confine the pass band from 480 to 763 mn. It simultaneously monitors 6000 stars brighter than 12th magnitude within a single star <span class="hlt">field</span> in the galactic plane. When the data are folded and phased to discover low amplitude transits, the relative precision of one-hour samples is about 1 part per thousand (10 x l0(exp -3)) for many of the brighter stars. This precision is sufficient to find jovian-size planets orbiting solar-like stars, which have signal amplitudes from 5 to 30 x l0(exp -3) depending on the inflation of the planet and the size of the star. Based on the frequency of giant inner-planets discovered by Doppler-<span class="hlt">velocity</span> method, one or two planets should be detectable in a rich star <span class="hlt">field</span>. The goal of the <span class="hlt">observations</span> is to obtain the sizes of giant extrasolar planets in short-period orbits and to combine these with masses determined from Doppler <span class="hlt">velocity</span> measurements to determine the densities of these planets. A further goal is to compare the measured planetary diameters with those predicted from theoretical models. From August 10 through September 30 of 1998, a forty nine square degree <span class="hlt">field</span> in the Cygnus constellation centered at RA and DEC of 19 hr 47 min, +36 deg 55 min was <span class="hlt">observed</span>. Useful data were obtained on twenty-nine nights. Nearly fifty stars showed some evidence of transits with periods between 0.3 and 8 days. Most had amplitudes too large to be associated with planetary transits. However, several stars showed low amplitude transits. The data for several transits of each of these two stars have been folded and been folded into 30 minute periods. Only Cygl433 shows any evidence of a flattened bottom that is expected when a small object</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19900061854&hterms=1052&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3D%2526%25231052','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900061854&hterms=1052&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3D%2526%25231052"><span><span class="hlt">Velocity</span> mapping and models of the elliptical galaxies NGC 720, NGC 1052, and NGC 4697</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Binney, J. J.; Davies, Roger L.; Illingworth, Garth D.</p> <p>1990-01-01</p> <p>CCD surface photometry and extensive long-slit spectroscopy are used to construct detailed models of the flattened ellipticals NGC 720, 1052, and 4697. The models are combined with the Jeans equations to yield predicted <span class="hlt">fields</span> of line-of-sight <span class="hlt">velocity</span> dispersion and streaming <span class="hlt">velocity</span>. By comparing these <span class="hlt">fields</span> with <span class="hlt">observed</span> <span class="hlt">velocities</span>, it is concluded that none of these systems can have isotropic <span class="hlt">velocity</span> dispersion tensors, and diminishing the assumed inclination of any given galaxy tends to decrease the line-of-sight <span class="hlt">velocity</span> dispersion and, counterintuitively, to increase the line-of-sight rotation speeds. The ratio of the line-of-sight <span class="hlt">velocity</span> dispersion along the minor axis to that along the major axis is found to be a sensitive diagnostic of the importance of a third integral for the galaxy's structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24381284','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24381284"><span>Repeating firing <span class="hlt">fields</span> of CA1 neurons shift forward in response to increasing angular <span class="hlt">velocity</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cowen, Stephen L; Nitz, Douglas A</p> <p>2014-01-01</p> <p>Self-motion information influences spatially-specific firing patterns exhibited by hippocampal neurons. Moreover, these firing patterns can repeat across similar subsegments of an environment, provided that there is similarity of path shape and head orientations across subsegments. The influence of self-motion variables on repeating <span class="hlt">fields</span> remains to be determined. To investigate the role of path shape and angular rotation on hippocampal activity, we recorded the activity of CA1 neurons from rats trained to run on spiral-shaped tracks. During inbound traversals of circular-spiral tracks, angular <span class="hlt">velocity</span> increases continuously. Under this condition, most neurons (74%) exhibited repeating <span class="hlt">fields</span> across at least three adjacent loops. Of these neurons, 86% exhibited forward shifts in the angles of <span class="hlt">field</span> centers relative to centers on preceding loops. Shifts were absent on squared-spiral tracks, minimal and less reliable on concentric-circle tracks, and absent on outward-bound runs on circular-spiral tracks. However, outward-bound runs on the circular-spiral track in the dark were associated with backward shifts. Together, the most parsimonious interpretation of the results is that continuous increases or decreases in angular <span class="hlt">velocity</span> are particularly effective at shifting the center of mass of repeating <span class="hlt">fields</span>, although it is also possible that a nonlinear integration of step counts contributes to the shift. Furthermore, the unexpected absence of <span class="hlt">field</span> shifts during outward journeys in light (but not darkness) suggests visual cues around the goal location anchored the map of space to an allocentric reference frame.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27315449','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27315449"><span>The effect of spatially varying <span class="hlt">velocity</span> <span class="hlt">field</span> on the transport of radioactivity in a porous medium.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sen, Soubhadra; Srinivas, C V; Baskaran, R; Venkatraman, B</p> <p>2016-10-01</p> <p>In the event of an accidental leak of the immobilized nuclear waste from an underground repository, it may come in contact of the flow of underground water and start migrating. Depending on the nature of the geological medium, the flow <span class="hlt">velocity</span> of water may vary spatially. Here, we report a numerical study on the migration of radioactivity due to a space dependent flow <span class="hlt">field</span>. For a detailed analysis, seven different types of <span class="hlt">velocity</span> profiles are considered and the corresponding concentrations are compared. Copyright © 2016 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhyA..491..946L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhyA..491..946L"><span>Update schemes of multi-<span class="hlt">velocity</span> floor <span class="hlt">field</span> cellular automaton for pedestrian dynamics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Luo, Lin; Fu, Zhijian; Cheng, Han; Yang, Lizhong</p> <p>2018-02-01</p> <p>Modeling pedestrian movement is an interesting problem both in statistical physics and in computational physics. Update schemes of cellular automaton (CA) models for pedestrian dynamics govern the schedule of pedestrian movement. Usually, different update schemes make the models behave in different ways, which should be carefully recalibrated. Thus, in this paper, we investigated the influence of four different update schemes, namely parallel/synchronous scheme, random scheme, order-sequential scheme and shuffled scheme, on pedestrian dynamics. The multi-<span class="hlt">velocity</span> floor <span class="hlt">field</span> cellular automaton (FFCA) considering the changes of pedestrians' moving properties along walking paths and heterogeneity of pedestrians' walking abilities was used. As for parallel scheme only, the collisions detection and resolution should be considered, resulting in a great difference from any other update schemes. For pedestrian evacuation, the evacuation time is enlarged, and the difference in pedestrians' walking abilities is better reflected, under parallel scheme. In face of a bottleneck, for example a exit, using a parallel scheme leads to a longer congestion period and a more dispersive density distribution. The exit flow and the space-time distribution of density and <span class="hlt">velocity</span> have significant discrepancies under four different update schemes when we simulate pedestrian flow with high desired <span class="hlt">velocity</span>. Update schemes may have no influence on pedestrians in simulation to create tendency to follow others, but sequential and shuffled update scheme may enhance the effect of pedestrians' familiarity with environments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011JGRA..11611202G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011JGRA..11611202G"><span>Dynamic subauroral ionospheric electric <span class="hlt">fields</span> <span class="hlt">observed</span> by the Falkland Islands radar during the course of a geomagnetic storm</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Grocott, A.; Milan, S. E.; Baker, J. B. H.; Freeman, M. P.; Lester, M.; Yeoman, T. K.</p> <p>2011-11-01</p> <p>We present an analysis of ionospheric electric <span class="hlt">field</span> data <span class="hlt">observed</span> during a geomagnetic storm by the recently deployed HF radar located on the Falkland Islands. On 3 August 2010 at ˜1800 UT evidence of the onset of a geomagnetic storm was <span class="hlt">observed</span> in ground magnetometer data in the form of a decrease in the Sym-H index of ˜100 nT. The main phase of the storm was <span class="hlt">observed</span> to last ˜24 hours before a gradual recovery lasting ˜3 days. On 4 August, during the peak magnetic disturbance of the storm, a high <span class="hlt">velocity</span> (>1000 m s-1) channel of ionospheric plasma flow, which we interpret as a subauroral ion drift (SAID), located between 53° and 58° magnetic south and lasting ˜6.5 hours, was <span class="hlt">observed</span> by the Falkland Islands radar in the pre-midnight sector. Coincident flow data from the DMSP satellites and the magnetically near-conjugate northern hemisphere Blackstone HF radar reveal that the SAID was embedded within the broader subauroral polarization streams (SAPS). DMSP particle data indicate that the SAID location closely followed the equatorward edge of the auroral electron precipitation boundary, while remaining generally poleward of the equatorward boundary of the ion precipitation. The latitude of the SAID varied throughout the interval on similar timescales to variations in the interplanetary magnetic <span class="hlt">field</span> and auroral activity, while variations in its <span class="hlt">velocity</span> were more closely related to ring current dynamics. These results are consistent with SAID electric <span class="hlt">fields</span> being generated by localized charge separation in the partial ring current, but suggest that their location is more strongly governed by solar wind driving and associated large-scale magnetospheric dynamics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19905431','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19905431"><span>Influence of anisotropy on anomalous scaling of a passive scalar advected by the Navier-Stokes <span class="hlt">velocity</span> <span class="hlt">field</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jurcisinová, E; Jurcisin, M; Remecký, R</p> <p>2009-10-01</p> <p>The influence of weak uniaxial small-scale anisotropy on the stability of the scaling regime and on the anomalous scaling of the single-time structure functions of a passive scalar advected by the <span class="hlt">velocity</span> <span class="hlt">field</span> governed by the stochastic Navier-Stokes equation is investigated by the <span class="hlt">field</span> theoretic renormalization group and operator-product expansion within one-loop approximation of a perturbation theory. The explicit analytical expressions for coordinates of the corresponding fixed point of the renormalization-group equations as functions of anisotropy parameters are found, the stability of the three-dimensional Kolmogorov-like scaling regime is demonstrated, and the dependence of the borderline dimension d(c) is an element of (2,3] between stable and unstable scaling regimes is found as a function of the anisotropy parameters. The dependence of the turbulent Prandtl number on the anisotropy parameters is also briefly discussed. The influence of weak small-scale anisotropy on the anomalous scaling of the structure functions of a passive scalar <span class="hlt">field</span> is studied by the operator-product expansion and their explicit dependence on the anisotropy parameters is present. It is shown that the anomalous dimensions of the structure functions, which are the same (universal) for the Kraichnan model, for the model with finite time correlations of the <span class="hlt">velocity</span> <span class="hlt">field</span>, and for the model with the advection by the <span class="hlt">velocity</span> <span class="hlt">field</span> driven by the stochastic Navier-Stokes equation in the isotropic case, can be distinguished by the assumption of the presence of the small-scale anisotropy in the systems even within one-loop approximation. The corresponding comparison of the anisotropic anomalous dimensions for the present model with that obtained within the Kraichnan rapid-change model is done.</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('http://adsabs.harvard.edu/abs/2017EGUGA..19.7228S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.7228S"><span>Study on <span class="hlt">field</span>-aligned electrons with Cluster <span class="hlt">observation</span> in the Earth's cusp</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shi, Jiankui; Torkar, Klaus; Cheng, Zhengwei</p> <p>2017-04-01</p> <p>Cusp region is very important to the solar wind-magnetosphere coupling. The solar wind particles, through the cusp, can directly entry into the magnetosphere and ionosphere, and transport the mass, momentum and energy. The gyrating charged particles with <span class="hlt">field</span>-aligned <span class="hlt">velocity</span> are significant to perform the transportation. In this study, data from Cluster <span class="hlt">observation</span> are used to study the characteristics of <span class="hlt">field</span>-aligned electrons (FAE's) including the downward and the upward FAEs in the cusp. We select FAE event to do analysis. The durations of the FAE event covered a wide range from 6 to 475 seconds. The FAE's were found to occur very commonly in a circumpolar zone in the polar region and the MLT and ILAT distributions showed that most of the FAE events were <span class="hlt">observed</span> around the cusp (70-80°ILAT, 0900-1500MLT). With the FAE flux the contribution of the electrons to the <span class="hlt">Field</span>-Aligned Current (FAC) is estimated and the result shows that the FAE was the main carrier to the FAC in the cusp. The physical mechanisms of the FAE are analyzed, namely that the downward electrons were mainly from the solar wind and the upward electrons may originated from accelerated ionospheric up-flowing electrons or mirrored solar wind electrons. The energy transportation into the magnetosphere by the solar wind electrons through the cusp is also investigated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PPCF...59b5013P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PPCF...59b5013P"><span>Magnetic <span class="hlt">field</span> pitch angle and perpendicular <span class="hlt">velocity</span> measurements from multi-point time-delay estimation of poloidal correlation reflectometry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Prisiazhniuk, D.; Krämer-Flecken, A.; Conway, G. D.; Happel, T.; Lebschy, A.; Manz, P.; Nikolaeva, V.; Stroth, U.; the ASDEX Upgrade Team</p> <p>2017-02-01</p> <p>In fusion machines, turbulent eddies are expected to be aligned with the direction of the magnetic <span class="hlt">field</span> lines and to propagate in the perpendicular direction. Time delay measurements of density fluctuations can be used to calculate the magnetic <span class="hlt">field</span> pitch angle α and perpendicular <span class="hlt">velocity</span> {{v}\\bot} profiles. The method is applied to poloidal correlation reflectometry installed at ASDEX Upgrade and TEXTOR, which measure density fluctuations from poloidally and toroidally separated antennas. Validation of the method is achieved by comparing the perpendicular <span class="hlt">velocity</span> (composed of the E× B drift and the phase <span class="hlt">velocity</span> of turbulence {{v}\\bot}={{v}E× B}+{{v}\\text{ph}} ) with Doppler reflectometry measurements and with neoclassical {{v}E× B} calculations. An important condition for the application of the method is the presence of turbulence with a sufficiently long decorrelation time. It is shown that at the shear layer the decorrelation time is reduced, limiting the application of the method. The magnetic <span class="hlt">field</span> pitch angle measured by this method shows the expected dependence on the magnetic <span class="hlt">field</span>, plasma current and radial position. The profile of the pitch angle reproduces the expected shape and values. However, comparison with the equilibrium reconstruction code cliste suggests an additional inclination of turbulent eddies at the pedestal position (2-3°). This additional angle decreases towards the core and at the edge.</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('http://adsabs.harvard.edu/abs/2018PhyA..492.2154W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhyA..492.2154W"><span>A new car-following model for autonomous vehicles flow with mean expected <span class="hlt">velocity</span> <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wen-Xing, Zhu; Li-Dong, Zhang</p> <p>2018-02-01</p> <p>Due to the development of the modern scientific technology, autonomous vehicles may realize to connect with each other and share the information collected from each vehicle. An improved forward considering car-following model was proposed with mean expected <span class="hlt">velocity</span> <span class="hlt">field</span> to describe the autonomous vehicles flow behavior. The new model has three key parameters: adjustable sensitivity, strength factor and mean expected <span class="hlt">velocity</span> <span class="hlt">field</span> size. Two lemmas and one theorem were proven as criteria for judging the stability of homogeneousautonomous vehicles flow. Theoretical results show that the greater parameters means larger stability regions. A series of numerical simulations were carried out to check the stability and fundamental diagram of autonomous flow. From the numerical simulation results, the profiles, hysteresis loop and density waves of the autonomous vehicles flow were exhibited. The results show that with increased sensitivity, strength factor or <span class="hlt">field</span> size the traffic jam was suppressed effectively which are well in accordance with the theoretical results. Moreover, the fundamental diagrams corresponding to three parameters respectively were obtained. It demonstrates that these parameters play almost the same role on traffic flux: i.e. before the critical density the bigger parameter is, the greater flux is and after the critical density, the opposite tendency is. In general, the three parameters have a great influence on the stability and jam state of the autonomous vehicles flow.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19780018449','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19780018449"><span>Induced <span class="hlt">velocity</span> <span class="hlt">field</span> of a jet in a crossflow</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fearn, R. L.; Weston, R. P.</p> <p>1978-01-01</p> <p>An experimental investigation of a subsonic round jet exhausting perpendicularly from a flat plate into a subsonic crosswind of the same temperature was conducted. <span class="hlt">Velocity</span> and pressure measurements were made in planes perpendicular to the path of the jet for ratios of jet <span class="hlt">velocity</span> to crossflow <span class="hlt">velocity</span> ranging from 3 to 10. The results of these measurements are presented in tabular and graphical forms. A pair of diffuse contrarotating vortices is identified as a significant feature of the flow, and the characteristics of the vortices are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002EGSGA..27..616K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002EGSGA..27..616K"><span><span class="hlt">Velocities</span> of Auroral Coherent Echoes At 12 and 144 Mhz</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Koustov, A. V.; Danskin, D. W.; Makarevitch, R. A.; Uspensky, M. V.; Janhunen, P.; Nishitani, N.; Nozawa, N.; Lester, M.; Milan, S.</p> <p></p> <p>Two Doppler coherent radar systems are currently working at Hankasalmi, Finland, the STARE and CUTLASS radars operating at 144 MHz and 12 MHz, respectively. The STARE beam 3 is nearly co-located with the CUTLASS beam 5 providing an opportunity for echo <span class="hlt">velocity</span> comparison along the same direction but at significantly different radar frequencies. In this study we consider one event when STARE radar echoes are detected t the same ranges as CUTLASS radar echoes. The <span class="hlt">observations</span> are complemented by EISCAT measurements of the ionospheric electric <span class="hlt">field</span> and elec- tron density behavior at one range of 900 km. Two separate situations are studied; for the first one, CUTLASS <span class="hlt">observed</span> F-region echoes (including the range of the EIS- CAT measurements) while for the second one CUTLASS <span class="hlt">observed</span> E-region echoes. In both cases STARE E-region measurements were available. We show that F-region CUTLASS <span class="hlt">velocities</span> agree well with the convection component along the CUTLASS radar beam while STARE <span class="hlt">velocities</span> are sometimes smaller by a factor of 2-3. For the second case, STARE <span class="hlt">velocities</span> are found to be either smaller or larger than CUTLASS <span class="hlt">velocities</span>, depending on range. Plasma physics of E- and F-region irregularities is dis- cussed in attempt to explain inferred relationship between various <span class="hlt">velocities</span>. Special attention is paid to ionospheric refraction that is important for the detection of 12-MHz echoes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.A43G0363F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.A43G0363F"><span>Ship-based <span class="hlt">Observations</span> of Turbulence and Stratocumulus Cloud Microphysics in the SE Pacific Ocean from the VOCALS <span class="hlt">Field</span> Program</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fairall, C. W.; Williams, C.; Grachev, A. A.; Brewer, A.; Choukulkar, A.</p> <p>2013-12-01</p> <p>The VAMOS (VOCALS) <span class="hlt">field</span> program involved deployment of several measurement systems based on ships, land and aircraft over the SE Pacific Ocean. The NOAA Ship Ronald H. Brown was the primary platform for surface based measurements which included the High Resolution Doppler Lidar (HRDL) and the motion-stabilized 94-GHz cloud Doppler radar (W-band radar). In this paper, the data from the W-band radar will be used to study the turbulent and microphysical structure of the stratocumulus clouds prevalent in the region. The radar data consists of a 3 Hz time series of radar parameters (backscatter coefficient, mean Doppler shift, and Doppler width) at 175 range gates (25-m spacing). Several statistical methods to de-convolve the turbulent <span class="hlt">velocity</span> and gravitational settling <span class="hlt">velocity</span> are examined and an optimized algorithm is developed. 20 days of <span class="hlt">observations</span> are processed to examine in-cloud profiles of mean turbulent statistics (vertical <span class="hlt">velocity</span> variance, skewness, dissipation rate) in terms of surface fluxes and estimates of entrainment and cloudtop radiative cooling. The clean separation of turbulent and fall <span class="hlt">velocities</span> will allow us to compute time-averaged drizzle-drop size spectra within and below the cloud that are significantly superior to previous attempts with surface-based marine cloud radar <span class="hlt">observations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/fs/2013/3028/pdf/fs2013-3028.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/fs/2013/3028/pdf/fs2013-3028.pdf"><span>Visualizing flow <span class="hlt">fields</span> using acoustic Doppler current profilers and the <span class="hlt">Velocity</span> Mapping Toolbox</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Jackson, P. Ryan</p> <p>2013-01-01</p> <p>The purpose of this fact sheet is to provide examples of how the U.S. Geological Survey is using acoustic Doppler current profilers for much more than routine discharge measurements. These instruments are capable of mapping complex three-dimensional flow <span class="hlt">fields</span> within rivers, lakes, and estuaries. Using the <span class="hlt">Velocity</span> Mapping Toolbox to process the ADCP data allows detailed visualization of the data, providing valuable information for a range of studies and applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/8817953','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/8817953"><span>Estimating mechanical blood trauma in a centrifugal blood pump: laser Doppler anemometer measurements of the mean <span class="hlt">velocity</span> <span class="hlt">field</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pinotti, M; Paone, N</p> <p>1996-06-01</p> <p>A laser Doppler anemometer (LDA) was used to obtain the mean <span class="hlt">velocity</span> and the Reynolds stress <span class="hlt">fields</span> in the inner channels of a well-known centrifugal vaneless pump (Bio-pump). Effects of the excessive flow resistance against which an occlusive pump operates in some surgical situations, such as cardiopulmonary bypass, are illustrated. The <span class="hlt">velocity</span> vector <span class="hlt">field</span> obtained from LDA measurements reveals that the constraint-forced vortex provides pumping action in a restricted area in the core of the pump. In such situations, recirculating zones dominate the flow and consequently increase the damage to blood cells and raise the risk of thrombus formation in the device. Reynolds normal and shear stress <span class="hlt">fields</span> were obtained in the entry flow for the channel formed by two rotating cones to illustrate the effects of flow disturbances on the potential for blood cell damage.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ARA%26A..55..111H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ARA%26A..55..111H"><span><span class="hlt">Observing</span> Interstellar and Intergalactic Magnetic <span class="hlt">Fields</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Han, J. L.</p> <p>2017-08-01</p> <p><span class="hlt">Observational</span> results of interstellar and intergalactic magnetic <span class="hlt">fields</span> are reviewed, including the <span class="hlt">fields</span> in supernova remnants and loops, interstellar filaments and clouds, Hii regions and bubbles, the Milky Way and nearby galaxies, galaxy clusters, and the cosmic web. A variety of approaches are used to investigate these <span class="hlt">fields</span>. The orientations of magnetic <span class="hlt">fields</span> in interstellar filaments and molecular clouds are traced by polarized thermal dust emission and starlight polarization. The <span class="hlt">field</span> strengths and directions along the line of sight in dense clouds and cores are measured by Zeeman splitting of emission or absorption lines. The large-scale magnetic <span class="hlt">fields</span> in the Milky Way have been best probed by Faraday rotation measures of a large number of pulsars and extragalactic radio sources. The coherent Galactic magnetic <span class="hlt">fields</span> are found to follow the spiral arms and have their direction reversals in arms and interarm regions in the disk. The azimuthal <span class="hlt">fields</span> in the halo reverse their directions below and above the Galactic plane. The orientations of organized magnetic <span class="hlt">fields</span> in nearby galaxies have been <span class="hlt">observed</span> through polarized synchrotron emission. Magnetic <span class="hlt">fields</span> in the intracluster medium have been indicated by diffuse radio halos, polarized radio relics, and Faraday rotations of embedded radio galaxies and background sources. Sparse evidence for very weak magnetic <span class="hlt">fields</span> in the cosmic web is the detection of the faint radio bridge between the Coma cluster and A1367. Future <span class="hlt">observations</span> should aim at the 3D tomography of the large-scale coherent magnetic <span class="hlt">fields</span> in our Galaxy and nearby galaxies, a better description of intracluster <span class="hlt">field</span> properties, and firm detections of intergalactic magnetic <span class="hlt">fields</span> in the cosmic web.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NHESS..18....1T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NHESS..18....1T"><span>Exploiting LSPIV to assess debris-flow <span class="hlt">velocities</span> in the <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Theule, Joshua I.; Crema, Stefano; Marchi, Lorenzo; Cavalli, Marco; Comiti, Francesco</p> <p>2018-01-01</p> <p>The assessment of flow <span class="hlt">velocity</span> has a central role in quantitative analysis of debris flows, both for the characterization of the phenomenology of these processes and for the assessment of related hazards. Large-scale particle image velocimetry (LSPIV) can contribute to the assessment of surface <span class="hlt">velocity</span> of debris flows, provided that the specific features of these processes (e.g. fast stage variations and particles up to boulder size on the flow surface) are taken into account. Three debris-flow events, each of them consisting of several surges featuring different sediment concentrations, flow stages, and <span class="hlt">velocities</span>, have been analysed at the inlet of a sediment trap in a stream in the eastern Italian Alps (Gadria Creek). Free software has been employed for preliminary treatment (orthorectification and format conversion) of video-recorded images as well as for LSPIV application. Results show that LSPIV <span class="hlt">velocities</span> are consistent with manual measurements of the orthorectified imagery and with front <span class="hlt">velocity</span> measured from the hydrographs in a channel recorded approximately 70 m upstream of the sediment trap. Horizontal turbulence, computed as the standard deviation of the flow directions at a given cross section for a given surge, proved to be correlated with surface <span class="hlt">velocity</span> and with visually estimated sediment concentration. The study demonstrates the effectiveness of LSPIV in the assessment of surface <span class="hlt">velocity</span> of debris flows and permit the most crucial aspects to be identified in order to improve the accuracy of debris-flow <span class="hlt">velocity</span> measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910031880&hterms=hydra&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dhydra','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910031880&hterms=hydra&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dhydra"><span>Potential, <span class="hlt">velocity</span>, and density <span class="hlt">fields</span> from redshift-distance samples: Application - Cosmography within 6000 kilometers per second</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bertschinger, Edmund; Dekel, Avishai; Faber, Sandra M.; Dressler, Alan; Burstein, David</p> <p>1990-01-01</p> <p>A potential flow reconstruction algorithm has been applied to the real universe to reconstruct the three-dimensional potential, <span class="hlt">velocity</span>, and mass density <span class="hlt">fields</span> smoothed on large scales. The results are shown as maps of these <span class="hlt">fields</span>, revealing the three-dimensional structure within 6000 km/s distance from the Local Group. The dominant structure is an extended deep potential well in the Hydra-Centaurus region, stretching across the Galactic plane toward Pavo, broadly confirming the Great Attractor (GA) model of Lynden-Bell et al. (1988). The Local Supercluster appears to be an extended ridge on the near flank of the GA, proceeding through the Virgo Southern Extension to the Virgo and Ursa Major clusters. The Virgo cluster and the Local Group are both falling toward the bottom of the GA potential well with peculiar <span class="hlt">velocities</span> of 658 + or - 121 km/s and 565 + or - 125 km/s, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..16..691O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..16..691O"><span><span class="hlt">Velocity</span> profiles, Reynolds stresses and bed roughness from an autonomous <span class="hlt">field</span> deployed Acoustic Doppler <span class="hlt">Velocity</span> Profiler in a mixed sediment tidal estuary</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>O'Boyle, Louise; Thorne, Peter; Cooke, Richard; Cohbed Team</p> <p>2014-05-01</p> <p>Estuaries are among some of the most important global landscapes in terms of population density, ecology and economy. Understanding the dynamics of these natural mixed sediment environments is of particular interest amid growing concerns over sea level rise, climate variations and estuarine response to these changes. Many predictors exist for bed form formation and sand transport in sandy coastal zones; however less work has been published on mixed sediments. This paper details a <span class="hlt">field</span> study which forms part of the COHBED project aiming to increase understanding of bed forms in a biotic mixed sediment estuarine environment. The study was carried out in the Dee Estuary, in the eastern Irish Sea between England and Wales from the 21st May to 4th June 2013. A state of the art instrumentation frame, known as SEDbed, was deployed at three sites of differing sediment properties and biological makeup within the intertidal zone of the estuary. The SEDbed deployment consisted of a suite of optical and acoustic instrumentation, including an Acoustic Doppler <span class="hlt">Velocity</span> Profiler (ADVP), Acoustic Doppler Velocimeter (ADV) and a three dimensional acoustic ripple profiler, 3D-ARP. Supplementary <span class="hlt">field</span> samples and measurements were recorded alongside the frame during each deployment. This paper focuses on the use of new technological developments for the investigation of sediment dynamics. The hydrodynamics at each of the deployment sites are presented including centimetre resolution <span class="hlt">velocity</span> profiles in the near bed region of the water column, obtained from the ADVP, which is presently the only autonomous <span class="hlt">field</span> deployed coherent Doppler profiler . Based on these high resolution profiles variations in frictional <span class="hlt">velocity</span>, bed shear stress and roughness length are calculated. Comparisons are made with theoretical models and with Reynolds stress values obtained from ADV data at a single point within the ADVP profile and from ADVP data itself. Predictions of bed roughness at each</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('http://adsabs.harvard.edu/abs/2011ApPhB.103..271K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011ApPhB.103..271K"><span>Simultaneous planar measurements of soot structure and <span class="hlt">velocity</span> <span class="hlt">fields</span> in a turbulent lifted jet flame at 3 kHz</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Köhler, M.; Boxx, I.; Geigle, K. P.; Meier, W.</p> <p>2011-05-01</p> <p>We describe a newly developed combustion diagnostic for the simultaneous planar imaging of soot structure and <span class="hlt">velocity</span> <span class="hlt">fields</span> in a highly sooting, lifted turbulent jet flame at 3000 frames per second, or two orders of magnitude faster than "conventional" laser imaging systems. This diagnostic uses short pulse duration (8 ns), frequency-doubled, diode-pumped solid state (DPSS) lasers to excite laser-induced incandescence (LII) at 3 kHz, which is then imaged onto a high framerate CMOS camera. A second (dual-cavity) DPSS laser and CMOS camera form the basis of a particle image <span class="hlt">velocity</span> (PIV) system used to acquire 2-component <span class="hlt">velocity</span> <span class="hlt">field</span> in the flame. The LII response curve (measured in a laminar propane diffusion flame) is presented and the combined diagnostics then applied in a heavily sooting lifted turbulent jet flame. The potential challenges and rewards of application of this combined imaging technique at high speeds are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016DDA....4730201M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016DDA....4730201M"><span>Predicting the <span class="hlt">Velocity</span> Dispersions of the Dwarf Satellite Galaxies of Andromeda</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McGaugh, Stacy S.</p> <p>2016-05-01</p> <p>Dwarf Spheroidal galaxies in the Local Group are the faintest and most diffuse stellar systems known. They exhibit large mass discrepancies, making them popular laboratories for studying the missing mass problem. The PANDAS survey of M31 revealed dozens of new examples of such dwarfs. As these systems were discovered, it was possible to use the <span class="hlt">observed</span> photometric properties to predict their stellar <span class="hlt">velocity</span> dispersions with the modified gravity theory MOND. These predictions, made in advance of the <span class="hlt">observations</span>, have since been largely confirmed. A unique feature of MOND is that a structurally identical dwarf will behave differently when it is or is not subject to the external <span class="hlt">field</span> of a massive host like Andromeda. The role of this "external <span class="hlt">field</span> effect" is critical in correctly predicting the <span class="hlt">velocity</span> dispersions of dwarfs that deviate from empirical scaling relations. With continued improvement in the <span class="hlt">observational</span> data, these systems could provide a test of the strong equivalence principle.</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('https://www.ncbi.nlm.nih.gov/pubmed/28868717','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28868717"><span>Estimating Mechanical Blood Trauma in a Centrifugal Blood Pump: Laser Doppler Anemometer Measurements of the Mean <span class="hlt">Velocity</span> <span class="hlt">Field</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pinotti, Marcos; Paone, Nicola</p> <p>1996-05-01</p> <p>A laser Doppler anemometer (LDA) was used to obtain the mean <span class="hlt">velocity</span> and the Reynolds stress <span class="hlt">fields</span> in the inner channels of a well-known centrifugal vaneless pump (Bio-pump). Effects of the excessive flow resistance against which an occlusive pump operates in some surgical situations, such as cardiopulmonary bypass, are illustrated. The <span class="hlt">velocity</span> vector <span class="hlt">field</span> obtained from LDA measurements reveals that the constraint-forced vortex provides pumping action in a restricted area in the core of the pump. In such situations, recirculating zones dominate the flow and consequently increase the damage to blood cells and raise the risk of thrombus formation in the device. Reynolds normal and shear stress <span class="hlt">fields</span> were obtained in the entry flow for the channel formed by two rotating cones to illustrate the effects of flow disturbances on the potential for blood cell damage. © 1996 International Society for Artificial Organs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70125433','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70125433"><span><span class="hlt">Field</span> <span class="hlt">observations</span> of swash zone flow patterns and 3D morphodynamics</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Puelo, Jack A.; Holland, K. Todd; Kooney, Timothy N.; Sallenger,, Asbury H.; Edge, Billy L.</p> <p>2001-01-01</p> <p>Rapid video measurements of foreshore morphology and <span class="hlt">velocity</span> were collected at Duck, NC in 1997 to investigate sediment transport processes in the swash zone. Estimates of foreshore evolution over a roughly 30 m cross-shore by 80 m alongshore study area were determined using a stereogrammetric technique. During the passage of a small storm (offshore wave heights increased from 1.4 to 2.5 m), the foreshore eroded nearly 40 cm in less than 4 hours. Dense, horizontal surface <span class="hlt">velocities</span> were measured over a sub-region (roughly 30 m by 40 m) of the study area using a new particle image velocimetry technique. This technique was able to quantify <span class="hlt">velocities</span> across the bore front approaching 5 m s–1 as well as the rapid <span class="hlt">velocities</span> in the very shallow backwash flows. The <span class="hlt">velocity</span> and foreshore topography measurements were used to test a three-dimensional energetics-based sediment transport model. Even though these data represent the most extensive and highly resolved swash measurements to date, the results showed that while the model could predict some of the qualitative trends in the <span class="hlt">observed</span> foreshore change, it was a poor predictor of the <span class="hlt">observed</span> magnitudes of foreshore change. Model — data comparisons differed by roughly an order of magnitude with <span class="hlt">observed</span> foreshore changes on the order of 10's of centimeters and model predictions on the order of meters. This poor comparison suggests that future models of swash-zone sediment transport may require the inclusion of other physical processes such as bore turbulence, fluid accelerations and skewness, infiltration/exfiltration, water depth variations, and variable friction factors (to name a few).</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('http://hdl.handle.net/2060/19900012290','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19900012290"><span>SR90, strontium shaped-charge 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>Wescott, Eugene M.; Stenbaek-Nielsen, Hans; Swift, Daniel W.; Valenzuela, Arnoldo; Rees, David</p> <p>1990-01-01</p> <p>In May 1986 an experiment was performed to test Alfven's critical ionization <span class="hlt">velocity</span> (CIV) effect in free space, using the first high explosive shaped charge with a conical liner of strontium metal. The release, made at 540 km altitude at dawn twilight, was aimed at 48 deg to B. The background electron density was 1.5 x 10(exp 4) cu cm. A faint <span class="hlt">field</span>-aligned Sr(+) ion streak with tip <span class="hlt">velocity</span> of 2.6 km/s was <span class="hlt">observed</span> from two optical sites. Using two calibration methods, it was calculated that between 4.5 x 10(exp 20) and 2 x 10(exp 21) ions were visible. An ionization time constant of 1920 s was calculated for Sr from the solar UV spectrum and ionization cross section which combined with a computer simulation of the injection predicts 1.7 x 10(exp 21) solar UV ions in the low-<span class="hlt">velocity</span> part of the ion streak. Thus all the <span class="hlt">observed</span> ions are from solar UV ionization of the slow (less than critical) <span class="hlt">velocity</span> portion of the neutral jet. The <span class="hlt">observed</span> neutral Sr <span class="hlt">velocity</span> distribution and computer simulations indicate that 2 x 10(exp 21) solar UV ions would have been created from the fast (greater than critical) part of the jet. They would have been more diffuse, and were not <span class="hlt">observed</span>. Using this fact it was estimated that any CIV ions created were less than 10(exp 21). It was concluded that future Sr CIV free space experiments should be conducted below the UV shadow height and in much larger background plasma density.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMSM51A2530B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMSM51A2530B"><span><span class="hlt">Velocity</span> Space Evolution of Dayside Reconnection Outflow</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Broll, J. M.; Fuselier, S. A.; Trattner, K. J.</p> <p>2015-12-01</p> <p>Magnetic reconnection is a universal phenomenon occurring when energy stored in a complicated magnetic <span class="hlt">field</span> topology is released into the surrounding plasma as the <span class="hlt">field</span> simplifies its configuration. At Earth's dayside magnetopause, reconnection is responsible for mass and energy input from the solar wind into the magnetosphere. We describe the evolution of the <span class="hlt">velocity</span>-space evolution of plasma outflow from a dayside magnetic reconnection region. We analyze Cluster magnetopause crossings between 1 and 10 Earth radii from the reconnection X-line predicted by the maximum magnetic shear model. The effects of nonadiabatic processes, such as deformation of the profile due to finite-gyroradius-induced pitch-angle scattering and wave-particle interactions, are described. We compare <span class="hlt">observations</span> and simulation results to describe the outflow evolution and infer the <span class="hlt">field</span>-aligned distance between an <span class="hlt">observation</span> and the reconnection site producing it.</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/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 gradient. 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 gradient <span class="hlt">fields</span> 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 <span class="hlt">field</span> and the tomography models which places today's <span class="hlt">observed</span> <span class="hlt">fields</span> 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> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/355506-new-test-chamber-measure-material-emissions-under-controlled-air-velocity','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/355506-new-test-chamber-measure-material-emissions-under-controlled-air-velocity"><span>A new test chamber to measure material emissions under controlled air <span class="hlt">velocity</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>Bortoli, M. de; Ghezzi, E.; Knoeppel, H.</p> <p>1999-05-15</p> <p>A new 20-L glass chamber for the determination of VOC emissions from construction materials and consumer products under controlled air <span class="hlt">velocity</span> and turbulence is described. Profiles of air <span class="hlt">velocity</span> and turbulence, obtained with precisely positioned hot wire anemometric probes, show that the <span class="hlt">velocity</span> <span class="hlt">field</span> is homogeneous and that air <span class="hlt">velocity</span> is tightly controlled by the fan rotation speed; this overcomes the problem of selecting representative positions to measure air <span class="hlt">velocity</span> above a test specimen. First tests on material emissions show that the influence of air <span class="hlt">velocity</span> on the emission rate of VOCs is negligible for sources limited by internal diffusionmore » and strong for sources limited by evaporation. In a <span class="hlt">velocity</span> interval from 0.15 to 0.30 m s{sup {minus}1}, an emission rate increase of 50% has been <span class="hlt">observed</span> for pure n-decane and 1,4-dichlorobenzene and of 30% for 1,2-propanediol from a water-based paint. In contrast, no measurable influence of turbulence could be <span class="hlt">observed</span> during vaporization of 1,4-dichlorobenzene within a 3-fold turbulence interval. Investigations still underway show that the chamber has a high recovery for the heavier VOC (TXIB), even at low concentrations.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008SPIE.6940E..1GS','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008SPIE.6940E..1GS"><span>Near-<span class="hlt">field</span> <span class="hlt">observation</span> platform</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schlemmer, Harry; Baeurle, Constantin; Vogel, Holger</p> <p>2008-04-01</p> <p>A miniaturized near-<span class="hlt">field</span> <span class="hlt">observation</span> platform is presented comprising a sensitive daylight camera and an uncooled micro-bolometer thermal imager each equipped with a wide angle lens. Both cameras are optimised for a range between a few meters and 200 m. The platform features a stabilised line of sight and can therefore be used also on a vehicle when it is in motion. The line of sight either can be directed manually or the platform can be used in a panoramic mode. The video output is connected to a control panel where algorithms for moving target indication or tracking can be applied in order to support the <span class="hlt">observer</span>. The near-<span class="hlt">field</span> platform also can be netted with the vehicle system and the signals can be utilised, e.g. to designate a new target to the main periscope or the weapon sight.</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('https://www.ncbi.nlm.nih.gov/pubmed/23919917','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23919917"><span>Easy monitoring of <span class="hlt">velocity</span> <span class="hlt">fields</span> in microfluidic devices using spatiotemporal image correlation spectroscopy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Travagliati, Marco; Girardo, Salvatore; Pisignano, Dario; Beltram, Fabio; Cecchini, Marco</p> <p>2013-09-03</p> <p>Spatiotemporal image correlation spectroscopy (STICS) is a simple and powerful technique, well established as a tool to probe protein dynamics in cells. Recently, its potential as a tool to map <span class="hlt">velocity</span> <span class="hlt">fields</span> in lab-on-a-chip systems was discussed. However, the lack of studies on its performance has prevented its use for microfluidics applications. Here, we systematically and quantitatively explore STICS microvelocimetry in microfluidic devices. We exploit a simple experimental setup, based on a standard bright-<span class="hlt">field</span> inverted microscope (no fluorescence required) and a high-fps camera, and apply STICS to map liquid flow in polydimethylsiloxane (PDMS) microchannels. Our data demonstrates optimal 2D velocimetry up to 10 mm/s flow and spatial resolution down to 5 μm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70182238','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70182238"><span>Electrical guidance efficiency of downstream-migrating juvenile Sea Lamprey decreases with increasing water <span class="hlt">velocity</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Miehls, Scott M.; Johnson, Nicholas; Haro, Alexander</p> <p>2017-01-01</p> <p>We tested the efficacy of a vertically oriented <span class="hlt">field</span> of pulsed direct current (VEPDC) created by an array of vertical electrodes for guiding downstream-moving juvenile Sea Lampreys Petromyzon marinus to a bypass channel in an artificial flume at water <span class="hlt">velocities</span> of 10–50 cm/s. Sea Lampreys were more likely to be captured in the bypass channel than in other sections of the flume regardless of electric <span class="hlt">field</span> status (on or off) or water <span class="hlt">velocity</span>. Additionally, Sea Lampreys were more likely to be captured in the bypass channel when the VEPDC was active; however, an interaction between the effects of VEPDC and water <span class="hlt">velocity</span> was <span class="hlt">observed</span>, as the likelihood of capture decreased with increases in water <span class="hlt">velocity</span>. The distribution of Sea Lampreys shifted from right to left across the width of the flume toward the bypass channel when the VEPDC was active at water <span class="hlt">velocities</span> less than 25 cm/s. The VEPDC appeared to have no effect on Sea Lamprey distribution in the flume at water <span class="hlt">velocities</span> greater than 25 cm/s. We also conducted separate tests to determine the threshold at which Sea Lampreys would become paralyzed. Individuals were paralyzed at a mean power density of 37.0 µW/cm3. Future research should investigate the ability of juvenile Sea Lampreys to detect electric <span class="hlt">fields</span> and their specific behavioral responses to electric <span class="hlt">field</span> characteristics so as to optimize the use of this technology as a nonphysical guidance tool across variable water <span class="hlt">velocities</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ExFl...57..154M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ExFl...57..154M"><span><span class="hlt">Velocity</span> <span class="hlt">field</span> measurements in the wake of a propeller model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mukund, R.; Kumar, A. Chandan</p> <p>2016-10-01</p> <p>Turboprop configurations are being revisited for the modern-day regional transport aircrafts for their fuel efficiency. The use of laminar flow wings is an effort in this direction. One way to further improve their efficiency is by optimizing the flow over the wing in the propeller wake. Previous studies have focused on improving the gross aerodynamic characteristics of the wing. It is known that the propeller slipstream causes early transition of the boundary layer on the wing. However, an optimized design of the propeller and wing combination could delay this transition and decrease the skin friction drag. Such a wing design would require the detailed knowledge of the development of the slipstream in isolated conditions. There are very few studies in the literature addressing the requirements of transport aircraft having six-bladed propeller and cruising at a high propeller advance ratio. Low-speed wind tunnel experiments have been conducted on a powered propeller model in isolated conditions, measuring the <span class="hlt">velocity</span> <span class="hlt">field</span> in the vertical plane behind the propeller using two-component hot-wire anemometry. The data obtained clearly resolved the mean <span class="hlt">velocity</span>, the turbulence, the ensemble phase averages and the structure and development of the tip vortex. The turbulence in the slipstream showed that transition could be close to the leading edge of the wing, making it a fine case for optimization. The development of the wake with distance shows some interesting flow features, and the data are valuable for flow computation and optimization.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006PhDT........12F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006PhDT........12F"><span>Galactic disk dynamical tracers: Open clusters and the local Milky Way rotation curve and <span class="hlt">velocity</span> <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Frinchaboy, Peter Michael, III</p> <p></p> <p>Establishing the rotation curve of the Milky Way is one of the fundamental contributions needed to understand the Galaxy and its mass distribution. We have undertaken a systematic spectroscopic survey of open star clusters which can serve as tracers of Galactic disk dynamics. We report on our initial sample of 67 clusters for which the Hydra multi-fiber spectrographs on the WIYN and Blanco telescopes have delivered ~1-2 km s -1 radial <span class="hlt">velocities</span> (RVs) of many dozens of stars in the <span class="hlt">fields</span> of each cluster, which are used to derive cluster membership and bulk cluster kinematics when combined with Tycho-2 proper motions. The clusters selected for study have a broad spatial distribution in order to be sensitive to the disk <span class="hlt">velocity</span> <span class="hlt">field</span> in all Galactic quadrants and across a Galactocentric radius range as much as 3.0 kpc from the solar circle. Through analysis of the cluster sample, we find (1) the rotation <span class="hlt">velocity</span> of the Local Standard of Rest (LSR) is [Special characters omitted.] km s -1 , (2 ) the local rotation curve is declining with radius having a slope of -9.1 km s -1 kpc -1 , (3) we find (using R 0 = 8.5 kpc) the following Galactic parameters: A = 17.0 km s -1 kpc -1 and B = -8.9 km s -1 kpc -1 , which using a flat rotation curve and our determined values for the rotation <span class="hlt">velocity</span> of the LSR yields a Galaxy mass within 1.5 R 0 of M = 1.4 ± 0.2 × 10 11 [Spe cial characters omitted.] and a M/L of 9 [Special characters omitted.] . We also explore the distribution of the local <span class="hlt">velocity</span> <span class="hlt">field</span> and find evidence for non- circular motion due to the spiral arms. Additionally, a number of outer disk ( R gc > 12 kpc) open clusters, including Be29 and Sa1, are studied that have potentially critical leverage on radial, age and metallicity gradients in the outer Galactic disk. We find that the measured kinematics of Sa1 and Be29 are consistent with being associated with the Galactic anticenter stellar structure (GASS; or Monoceros stream), which points to a possible</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20110022982&hterms=CAPS&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DCAPS','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20110022982&hterms=CAPS&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DCAPS"><span>Background and Pickup Ion <span class="hlt">Velocity</span> Distribution Dynamics in Titan's Plasma Environment: 3D Hybrid Simulation and Comparison with CAPS T9 <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>Lipatov, A. S.; Sittler, E. C., Jr.; Hartle, R. E.; Cooper, J. F.; Simpson, D. G.</p> <p>2011-01-01</p> <p>In this report we discuss the ion <span class="hlt">velocity</span> distribution dynamics from the 3D hybrid simulation. In our model the background, pickup, and ionospheric ions are considered as a particles, whereas the electrons are described as a fluid. Inhomogeneous photoionization, electron-impact ionization and charge exchange are included in our model. We also take into account the collisions between the ions and neutrals. The current simulation shows that mass loading by pickup ions H(+); H2(+), CH4(+) and N2(+) is stronger than in the previous simulations when O+ ions are introduced into the background plasma. In our hybrid simulations we use Chamberlain profiles for the atmospheric components. We also include a simple ionosphere model with average mass M = 28 amu ions that were generated inside the ionosphere. The moon is considered as a weakly conducting body. Special attention will be paid to comparing the simulated pickup ion <span class="hlt">velocity</span> distribution with CAPS T9 <span class="hlt">observations</span>. Our simulation shows an asymmetry of the ion density distribution and the magnetic <span class="hlt">field</span>, including the formation of the Alfve n wing-like structures. The simulation also shows that the ring-like <span class="hlt">velocity</span> distribution for pickup ions relaxes to a Maxwellian core and a shell-like halo.</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('http://adsabs.harvard.edu/abs/2014ExFl...55.1681M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014ExFl...55.1681M"><span>On the relationship between image intensity and <span class="hlt">velocity</span> in a turbulent boundary layer seeded with smoke particles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Melnick, M. Blake; Thurow, Brian S.</p> <p>2014-02-01</p> <p>Simultaneous particle image velocimetry (PIV) and flow visualization measurements were performed in a turbulent boundary layer in an effort to better quantify the relationship between the <span class="hlt">velocity</span> <span class="hlt">field</span> and the image intensity typically <span class="hlt">observed</span> in a classical flow visualization experiment. The freestream flow was lightly seeded with smoke particles to facilitate PIV measurements, whereas the boundary layer was densely seeded with smoke through an upstream slit in the wall to facilitate both PIV and classical flow visualization measurements at Reynolds numbers, Re θ , ranging from 2,100 to 8,600. Measurements were taken with and without the slit covered as well as with and without smoke injection. The addition of a narrow slit in the wall produces a minor modification of the nominal turbulent boundary layer profile whose effect is reduced with downstream distance. The presence of dense smoke in the boundary layer had a minimal effect on the <span class="hlt">observed</span> <span class="hlt">velocity</span> <span class="hlt">field</span> and the associated proper orthogonal decomposition (POD) modes. Analysis of instantaneous images shows that the edge of the turbulent boundary layer identified from flow visualization images generally matches the edge of the boundary layer determined from <span class="hlt">velocity</span> and vorticity. The correlation between <span class="hlt">velocity</span> deficit and smoke intensity was determined to be positive and relatively large (>0.7) indicating a moderate-to-strong relationship between the two. This notion was extended further through the use of a direct correlation approach and a complementary POD/linear stochastic estimation (LSE) approach to estimate the <span class="hlt">velocity</span> <span class="hlt">field</span> directly from flow visualization images. This exercise showed that, in many cases, <span class="hlt">velocity</span> <span class="hlt">fields</span> estimated from smoke intensity were similar to the actual <span class="hlt">velocity</span> <span class="hlt">fields</span>. The complementary POD/LSE approach proved better for these estimations, but not enough to suggest using this technique to approximate <span class="hlt">velocity</span> measurements from a smoke intensity image. Instead, the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1995JCoPh.117..146E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1995JCoPh.117..146E"><span>A New Algorithm with Plane Waves and Wavelets for Random <span class="hlt">Velocity</span> <span class="hlt">Fields</span> with Many Spatial Scales</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Elliott, Frank W.; Majda, Andrew J.</p> <p>1995-03-01</p> <p>A new Monte Carlo algorithm for constructing and sampling stationary isotropic Gaussian random <span class="hlt">fields</span> with power-law energy spectrum, infrared divergence, and fractal self-similar scaling is developed here. The theoretical basis for this algorithm involves the fact that such a random <span class="hlt">field</span> is well approximated by a superposition of random one-dimensional plane waves involving a fixed finite number of directions. In general each one-dimensional plane wave is the sum of a random shear layer and a random acoustical wave. These one-dimensional random plane waves are then simulated by a wavelet Monte Carlo method for a single space variable developed recently by the authors. The computational results reported in this paper demonstrate remarkable low variance and economical representation of such Gaussian random <span class="hlt">fields</span> through this new algorithm. In particular, the <span class="hlt">velocity</span> structure function for an imcorepressible isotropic Gaussian random <span class="hlt">field</span> in two space dimensions with the Kolmogoroff spectrum can be simulated accurately over 12 decades with only 100 realizations of the algorithm with the scaling exponent accurate to 1.1% and the constant prefactor accurate to 6%; in fact, the exponent of the <span class="hlt">velocity</span> structure function can be computed over 12 decades within 3.3% with only 10 realizations. Furthermore, only 46,592 active computational elements are utilized in each realization to achieve these results for 12 decades of scaling behavior.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930053862&hterms=Muzzle&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DMuzzle','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930053862&hterms=Muzzle&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DMuzzle"><span>Stepwise shockwave <span class="hlt">velocity</span> determinator</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Roth, Timothy E.; Beeson, Harold</p> <p>1992-01-01</p> <p>To provide an uncomplicated and inexpensive method for measuring the far-<span class="hlt">field</span> <span class="hlt">velocity</span> of a surface shockwave produced by an explosion, a stepwise shockwave <span class="hlt">velocity</span> determinator (SSVD) was developed. The <span class="hlt">velocity</span> determinator is constructed of readily available materials and works on the principle of breaking discrete sensors composed of aluminum foil contacts. The discrete sensors have an average breaking threshold of approximately 7 kPa. An incremental output step of 250 mV is created with each foil contact breakage and is logged by analog-to-digital instrumentation. <span class="hlt">Velocity</span> data obtained from the SSVD is within approximately 11 percent of the calculated surface shockwave <span class="hlt">velocity</span> of a muzzle blast from a 30.06 rifle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009ApJ...704.1721P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009ApJ...704.1721P"><span>Magnetic <span class="hlt">Field</span> Topology in Low-Mass Stars: Spectropolarimetric <span class="hlt">Observations</span> of M Dwarfs</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Phan-Bao, Ngoc; Lim, Jeremy; Donati, Jean-François; Johns-Krull, Christopher M.; Martín, Eduardo L.</p> <p>2009-10-01</p> <p>The magnetic <span class="hlt">field</span> topology plays an important role in the understanding of stellar magnetic activity. While it is widely accepted that the dynamo action present in low-mass partially convective stars (e.g., the Sun) results in predominantly toroidal magnetic flux, the <span class="hlt">field</span> topology in fully convective stars (masses below ~0.35 M sun) is still under debate. We report here our mapping of the magnetic <span class="hlt">field</span> topology of the M4 dwarf G 164-31 (or Gl 490B), which is expected to be fully convective, based on time series data collected from 20 hr of <span class="hlt">observations</span> spread over three successive nights with the ESPaDOnS spectropolarimeter. Our tomographic imaging technique applied to time series of rotationally modulated circularly polarized profiles reveals an axisymmetric large-scale poloidal magnetic <span class="hlt">field</span> on the M4 dwarf. We then apply a synthetic spectrum fitting technique for measuring the average magnetic flux on the star. The flux measured in G 164-31 is |Bf| = 3.2 ± 0.4 kG, which is significantly greater than the average value of 0.68 kG determined from the imaging technique. The difference indicates that a significant fraction of the stellar magnetic energy is stored in small-scale structures at the surface of G 164-31. Our Hα emission light curve shows evidence for rotational modulation suggesting the presence of localized structure in the chromosphere of this M dwarf. The radius of the M4 dwarf derived from the rotational period and the projected equatorial <span class="hlt">velocity</span> is at least 30% larger than that predicted from theoretical models. We argue that this discrepancy is likely primarily due to the young nature of G 164-31 rather than primarily due to magnetic <span class="hlt">field</span> effects, indicating that age is an important factor which should be considered in the interpretation of this <span class="hlt">observational</span> result. We also report here our polarimetric <span class="hlt">observations</span> of five other M dwarfs with spectral types from M0 to M4.5, three of them showing strong Zeeman signatures. Based on</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ISPAr42.3..289D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ISPAr42.3..289D"><span>Copernicus Big Data and Google Earth Engine for Glacier Surface <span class="hlt">Velocity</span> <span class="hlt">Field</span> Monitoring: Feasibility Demonstration on San Rafael and San Quintin Glaciers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Di Tullio, M.; Nocchi, F.; Camplani, A.; Emanuelli, N.; Nascetti, A.; Crespi, M.</p> <p>2018-04-01</p> <p>The glaciers are a natural global resource and one of the principal climate change indicator at global and local scale, being influenced by temperature and snow precipitation changes. Among the parameters used for glacier monitoring, the surface <span class="hlt">velocity</span> is a key element, since it is connected to glaciers changes (mass balance, hydro balance, glaciers stability, landscape erosion). The leading idea of this work is to continuously retrieve glaciers surface <span class="hlt">velocity</span> using free ESA Sentinel-1 SAR imagery and exploiting the potentialities of the Google Earth Engine (GEE) platform. GEE has been recently released by Google as a platform for petabyte-scale scientific analysis and visualization of geospatial datasets. The algorithm of SAR off-set tracking developed at the Geodesy and Geomatics Division of the University of Rome La Sapienza has been integrated in a cloud based platform that automatically processes large stacks of Sentinel-1 data to retrieve glacier surface <span class="hlt">velocity</span> <span class="hlt">field</span> time series. We processed about 600 Sentinel-1 image pairs to obtain a continuous time series of <span class="hlt">velocity</span> <span class="hlt">field</span> measurements over 3 years from January 2015 to January 2018 for two wide glaciers located in the Northern Patagonian Ice <span class="hlt">Field</span> (NPIF), the San Rafael and the San Quintin glaciers. Several results related to these relevant glaciers also validated with respect already available and renown software (i.e. ESA SNAP, CIAS) and with respect optical sensor measurements (i.e. LANDSAT8), highlight the potential of the Big Data analysis to automatically monitor glacier surface <span class="hlt">velocity</span> <span class="hlt">fields</span> at global scale, exploiting the synergy between GEE and Sentinel-1 imagery.</p> </li> <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('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4518465','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4518465"><span>Ghost peaks <span class="hlt">observed</span> after AP-MALDI experiment may disclose new ionization mechanism of matrix assisted hypersonic <span class="hlt">velocity</span> impact ionization</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Moskovets, Eugene</p> <p>2015-01-01</p> <p>RATIONALE Understanding the mechanisms of MALDI promises improvements in the sensitivity and specificity of many established applications in the <span class="hlt">field</span> of mass spectrometry. This paper reports a serendipitous <span class="hlt">observation</span> of a significant ion yield in a post-ionization experiment conducted after the sample has been removed from a standard atmospheric pressure (AP)-MALDI source. This post-ionization is interpreted in terms of collisions of microparticles moving with a hypersonic <span class="hlt">velocity</span> into a solid surface. Calculations show that the thermal energy released during such collisions is close to that absorbed by the top matrix layer in traditional MALDI. The microparticles, containing both the matrix and analytes, could be detached from a film produced inside the inlet capillary during the sample ablation and accelerated by the flow rushing through the capillary. These <span class="hlt">observations</span> contribute some new perspective to ion formation in both laser and laserless matrix-assisted ionization. METHODS An AP-MALDI ion source hyphenated with a three-stage high-pressure ion funnel system was utilized for peptide mass analysis. After the laser was turned off and MALDI sample was removed, ions were detected during a gradual reduction of the background pressure in the first funnel. The constant-rate pressure reduction led to the reproducible appearance of different singly- and doubly-charged peptide peaks in mass spectra taken a few seconds after the end of the MALDI analysis of a dried-droplet spot. RESULTS The ion yield as well as the mass range of ions <span class="hlt">observed</span> with a significant delay after a completion of the primary MALDI analysis depended primarily on the background pressure inside the first funnel. The production of ions in this post-ionization step was exclusively <span class="hlt">observed</span> during the pressure drop. A lower matrix background and significant increase in relative yield of double-protonated ions are reported. CONCLUSIONS The <span class="hlt">observations</span> were partially consistent with a model 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_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/2014A%26A...563A..93Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014A%26A...563A..93Y"><span>The power spectrum of solar convection flows from high-resolution <span class="hlt">observations</span> and 3D simulations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yelles Chaouche, L.; Moreno-Insertis, F.; Bonet, J. A.</p> <p>2014-03-01</p> <p>Context. Understanding solar surface magnetoconvection requires the study of the Fourier spectra of the <span class="hlt">velocity</span> <span class="hlt">fields</span>. Nowadays, <span class="hlt">observations</span> are available that resolve very small spatial scales, well into the subgranular range, almost reaching the scales routinely resolved in numerical magnetoconvection simulations. Comparison of numerical and <span class="hlt">observational</span> data at present can provide an assessment of the validity of the <span class="hlt">observational</span> proxies. Aims: Our aims are: (1) to obtain Fourier spectra for the photospheric <span class="hlt">velocity</span> <span class="hlt">fields</span> using the spectropolarimetric <span class="hlt">observations</span> with the highest spatial resolution so far (~120 km), thus reaching for the first time spatial scales well into the subgranular range; (2) to calculate corresponding Fourier spectra from realistic 3D numerical simulations of magnetoconvection and carry out a proper comparison with their <span class="hlt">observational</span> counterparts considering the residual instrumental degradation in the <span class="hlt">observational</span> data; and (3) to test the <span class="hlt">observational</span> proxies on the basis of the numerical data alone, by comparing the actual <span class="hlt">velocity</span> <span class="hlt">field</span> in the simulations with synthetic <span class="hlt">observations</span> obtained from the numerical boxes. Methods: (a) For the <span class="hlt">observations</span>, data from the SUNRISE/IMaX spectropolarimeter are used. (b) For the simulations, we use four series of runs obtained with the STAGGER code for different average signed vertical magnetic <span class="hlt">field</span> values (0, 50, 100, and 200 G). Spectral line profiles are synthesized from the numerical boxes for the same line <span class="hlt">observed</span> by IMaX (Fe I 5250.2 Å) and degraded to match the performance of the IMaX instrument. Proxies for the <span class="hlt">velocity</span> <span class="hlt">field</span> are obtained via Dopplergrams (vertical component) and local correlation tracking (LCT, for the horizontal component). Fourier power spectra are calculated and a comparison between the synthetic and <span class="hlt">observational</span> data sets carried out. (c) For the internal comparison of the numerical data, <span class="hlt">velocity</span> values on constant optical depth surfaces are used</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018FrPhy..13.7203D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018FrPhy..13.7203D"><span>Electron drift <span class="hlt">velocity</span> and mobility in graphene</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dong, Hai-Ming; Duan, Yi-Feng; Huang, Fei; Liu, Jin-Long</p> <p>2018-04-01</p> <p>We present a theoretical study of the electric transport properties of graphene-substrate systems. The drift <span class="hlt">velocity</span>, mobility, and temperature of the electrons are self-consistently determined using the Boltzmann equilibrium equations. It is revealed that the electronic transport exhibits a distinctly nonlinear behavior. A very high mobility is achieved with the increase of the electric <span class="hlt">fields</span> increase. The electron <span class="hlt">velocity</span> is not completely saturated with the increase of the electric <span class="hlt">field</span>. The temperature of the hot electrons depends quasi-linearly on the electric <span class="hlt">field</span>. In addition, we show that the electron <span class="hlt">velocity</span>, mobility, and electron temperature are sensitive to the electron density. These findings could be employed for the application of graphene for high-<span class="hlt">field</span> nano-electronic devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUFMSM43A1475H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUFMSM43A1475H"><span>Lunar Electric <span class="hlt">Fields</span>: <span class="hlt">Observations</span> and Implications</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Halekas, J. S.; Delory, G. T.; Stubbs, T. J.; Farrell, W. M.; Vondrak, R. R.</p> <p>2006-12-01</p> <p>Alhough the Moon is typically thought of as having a relatively dormant environment, it is in fact very electrically active. The lunar surface, not protected by any substantial atmosphere, is directly exposed to solar UV and X-rays as well as solar wind plasma and energetic particles. This creates a complex electrodynamic environment, with the surface typically charging positive in sunlight and negative in shadow, and surface potentials varying over orders of magnitude in response to changing solar illumination and plasma conditions. <span class="hlt">Observations</span> from the Apollo era and theoretical considerations strongly suggest that surface charging also drives dust electrification and horizontal and vertical dust transport. We present a survey of the lunar electric <span class="hlt">field</span> environment, utilizing both newly interpreted Lunar Prospector (LP) orbital <span class="hlt">observations</span> and older Apollo surface <span class="hlt">observations</span>, and comparing to theoretical predictions. We focus in particular on time periods when the most significant surface charging was <span class="hlt">observed</span> by LP - namely plasmasheet crossings (when the Moon is in the Earth's magnetosphere) and space weather events. During these time periods, kV-scale potentials are <span class="hlt">observed</span>, and enhanced surface electric <span class="hlt">fields</span> can be expected to drive significant horizontal and vertical dust transport. Both dust and electric <span class="hlt">fields</span> can have serious effects on habitability and operation of machinery, so understanding the coupled dust-plasma-electric <span class="hlt">field</span> system around the Moon is critically important for planning exploration efforts, in situ resource utilization, and scientific <span class="hlt">observations</span> on the lunar surface. Furthermore, from a pure science perspective, this represents an excellent opportunity to study fundamental surface-plasma interactions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28616742','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28616742"><span>Wind-induced flow <span class="hlt">velocity</span> effects on nutrient concentrations at Eastern Bay of Lake Taihu, China.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jalil, Abdul; Li, Yiping; Du, Wei; Wang, Jianwei; Gao, Xiaomeng; Wang, Wencai; Acharya, Kumud</p> <p>2017-07-01</p> <p>Shallow lakes are highly sensitive to respond internal nutrient loading due to wind-induced flow <span class="hlt">velocity</span> effects. Wind-induced flow <span class="hlt">velocity</span> effects on nutrient suspension were investigated at a long narrow bay of large shallow Lake Taihu, the third largest freshwater lake in China. Wind-induced reverse/compensation flow and consistent flow <span class="hlt">field</span> probabilities at vertical column of the water were measured. The probabilities between the wind <span class="hlt">field</span> and the flow <span class="hlt">velocities</span> provided a strong correlation at the surface (80.6%) and the bottom (65.1%) layers of water profile. Vertical flow <span class="hlt">velocity</span> profile analysis provided the evidence of delay response time to wind <span class="hlt">field</span> at the bottom layer of lake water. Strong wind <span class="hlt">field</span> generated by the west (W) and west-north-west (WNW) winds produced displaced water movements in opposite directions to the prevailing flow <span class="hlt">field</span>. An exponential correlation was <span class="hlt">observed</span> between the current <span class="hlt">velocities</span> of the surface and the bottom layers while considering wind speed as a control factor. A linear model was developed to correlate the wind <span class="hlt">field</span>-induced flow <span class="hlt">velocity</span> impacts on nutrient concentration at the surface and bottom layers. Results showed that dominant wind directions (ENE, E, and ESE) had a maximum nutrient resuspension contribution (nutrient resuspension potential) of 34.7 and 43.6% at the surface and the bottom profile layers, respectively. Total suspended solids (TSS), total nitrogen (TN), and total phosphorus (TP) average concentrations were 6.38, 1.5, and 0.03 mg/L during our <span class="hlt">field</span> experiment at Eastern Bay of Lake Taihu. Overall, wind-induced low-to-moderate hydrodynamic disturbances contributed more in nutrient resuspension at Eastern Bay of Lake Taihu. The present study can be used to understand the linkage between wind-induced flow <span class="hlt">velocities</span> and nutrient concentrations for shallow lakes (with uniform morphology and deep margins) water quality management and to develop further models.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70155866','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70155866"><span><span class="hlt">Velocity</span> bias induced by flow patterns around ADCPs and associated deployment platforms</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Mueller, David S.</p> <p>2015-01-01</p> <p><span class="hlt">Velocity</span> measurements near the Acoustic Doppler Current Profiler (ADCP) are important for mapping surface currents, measuring <span class="hlt">velocity</span> and discharge in shallow streams, and providing accurate estimates of discharge in the top unmeasured portion of the water column. Improvements to ADCP performance permit measurement of <span class="hlt">velocities</span> much closer (5 cm) to the transducer than has been possible in the past (25 cm). <span class="hlt">Velocity</span> profiles collected by the U.S. Geological Survey (USGS) with a 1200 kHz Rio Grande Zedhead ADCP in 2002 showed a negative bias in measured <span class="hlt">velocities</span> near the transducers. On the basis of these results, the USGS initiated a study combining <span class="hlt">field</span>, laboratory, and numerical modeling data to assess the effect of flow patterns caused by flow around the ADCP and deployment platforms on <span class="hlt">velocities</span> measured near the transducers. This ongoing study has shown that the negative bias <span class="hlt">observed</span> in the <span class="hlt">field</span> is due to the flow pattern around the ADCP. The flow pattern around an ADCP violates the basic assumption of flow homogeneity required for an accurate three-dimensional <span class="hlt">velocity</span> solution. Results, to date (2014), have indicated <span class="hlt">velocity</span> biases within the measurable profile, due to flow disturbance, for the TRDI 1200 kHz Rio Grande Zedhead and the SonTek RiverSurveyor M9 ADCPs. The flow speed past the ADCP, the mount and the deployment platform have also been shown to play an important role in the magnitude and extent of the <span class="hlt">velocity</span> bias.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.S21A2668A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.S21A2668A"><span>Are seismic <span class="hlt">velocity</span> time-lapse changes due to fluid substitution or matrix dissolution? A CO2 sequestration study at Pohokura <span class="hlt">Field</span>, New Zealand</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Adam, L.; Sim, C. Y.; Macfarlane, J.; van Wijk, K.; Shragge, J. C.; Higgs, K.</p> <p>2015-12-01</p> <p>Time-lapse seismic signatures can be used to quantify fluid saturation and pressure changes in a reservoir undergoing CO2 sequestration. However, the injection of CO2 acidifies the water, which may dissolve and/or precipitate minerals. Understanding the impact on the rock frame from <span class="hlt">field</span> seismic time-lapse changes remains an outstanding challenge. Here, we study the effects of carbonate-CO2-water reactions on the physical and elastic properties of rock samples with variable volumes of carbonate cementation. The effects of fluid substitution alone (brine to CO2) and those due to the combination of fluid substitution and mineral dissolution on time-lapse seismic signatures are studied by combining laboratory data, geophysical well-log data and 1-D seismic modeling. Nine rocks from Pohokura <span class="hlt">Field</span> (New Zealand) are reacted with carbonic acid. The elastic properties are measured using a high-density laser-ultrasonic setup. We <span class="hlt">observe</span> that P-wave <span class="hlt">velocity</span> changes up to -19% and correlate with sandstone grain size. Coarse-grained sandstones show greater changes in elastic wave <span class="hlt">velocities</span> due to dissolution than fine-grained sandstones. To put this in perspective, this <span class="hlt">velocity</span> change is comparable to the effect of fluid substitution from brine to CO2. This can potentially create an ambiguity in the interpretation of the physical processes responsible for time-lapse signatures in a CO2injection scenario. The laboratory information is applied onto well-log data to model changes in elastic properties of sandstones at the well-log scale. Well-logs and core petrographic analyses are used to find an elastic model that best describes the <span class="hlt">observed</span> elastic waves <span class="hlt">velocities</span> in the cemented reservoir sandstones. The Constant-cement rock physics model is found to predict the elastic behaviour of the cemented sandstones. A possible late-time sequestration scenario is that both mineral dissolution and fluid substitution occur in the reservoir. 1-D synthetic seismograms show that</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1988STIN...8924566C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1988STIN...8924566C"><span><span class="hlt">Velocity</span> and scalar <span class="hlt">fields</span> of turbulent premixed flame in stagnation flow</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cho, P.; Law, C. K.; Cheng, R. K.; Shepherd, I. G.</p> <p>1988-08-01</p> <p>Detailed experimental measurements of the scalar and <span class="hlt">velocity</span> statistics of premixed methane/air flames stabilized by a stagnation plant are reported. Conditioned and unconditioned <span class="hlt">velocity</span> of two components and the reaction progress variables are measured by using a two-component laser Doppler velocimetry techniques and Mie scattering techniques, respectively. Experimental conditions cover equivalence ratios of 0.9 and 1.0, incident turbulence intensities of 0.3 to 0.45 m/s, and global stretch rates of 100 to 150 sec sup minus 1. The experimental results are analyzed in the context of the Bray-Moss-Libby flamelet model of these flames. The results indicate that there is no turbulence production within the turbulent flame brush and the second and third order turbulent transport terms are reduced to functions of the difference between the conditioned mean <span class="hlt">velocity</span>. The result of normalization of these relative <span class="hlt">velocities</span> by the respective <span class="hlt">velocity</span> increase across laminar flames suggest that the mean unconditioned <span class="hlt">velocity</span> profiles are self-similar.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvD..97f6020H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvD..97f6020H"><span>Holographic butterfly <span class="hlt">velocities</span> in brane geometry and Einstein-Gauss-Bonnet gravity with matters</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Huang, Wung-Hong</p> <p>2018-03-01</p> <p>In the first part of the paper we generalize the butterfly <span class="hlt">velocity</span> formula to anisotropic spacetime. We apply the formula to evaluate the butterfly <span class="hlt">velocities</span> in M-branes, D-branes, and strings backgrounds. We show that the butterfly <span class="hlt">velocities</span> in M2-branes, M5-branes and the intersection M 2 ⊥ M 5 equal to those in fundamental strings, D4-branes and the intersection F 1 ⊥ D 4 backgrounds, respectively. These <span class="hlt">observations</span> lead us to conjecture that the butterfly <span class="hlt">velocity</span> is generally invariant under a double-dimensional reduction. In the second part of the paper, we study the butterfly <span class="hlt">velocity</span> for Einstein-Gauss-Bonnet gravity with arbitrary matter <span class="hlt">fields</span>. A general formula is obtained. We use this formula to compute the butterfly <span class="hlt">velocities</span> in different backgrounds and discuss the associated properties.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19790008682','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19790008682"><span>Study of mean- and turbulent-<span class="hlt">velocity</span> <span class="hlt">fields</span> in a large-scale turbine-vane passage</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bailey, D. A.</p> <p>1979-01-01</p> <p>Laser-Doppler velocimetry, and to a lesser extent hot-wire anemometry, were employed to measure three components of the mean <span class="hlt">velocity</span> and the six turbulent stresses at four planes within the turbine inlet-guide-vane passage. One variation in the turbulent inlet boundary layer thickness and one variation in the blade aspect ratio (span/axial chord) were studied. A longitudinal vortex (passage vortex) was clearly identified in the exit plane of the passage for the three test cases. The maximum turbulence intensities within the longitudinal vortex were found to be on the order of 2 to 4 percent, with large regions appearing nonturbulent. Because a turbulent wall boundary layer was the source of vorticity that produced the passage vortex, these low turbulence levels were not anticipated. For the three test cases studied, the lateral <span class="hlt">velocity</span> <span class="hlt">field</span> extended significantly beyond the region of the longitudinal <span class="hlt">velocity</span> defect. Changing the inlet boundary layer thickness produced a difference in the location, the strength, and the extent of the passage vortex. Changing the aspect ratio of the blade passage had a measurable but less significant effect. The experiment was performed in a 210 mm pitch, 272 mm axial chord model in low speed wind tunnel at an inlet Mach number of 0.07.</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> <span class="hlt">fields</span> 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-gradient crust-mantle transition zone or a rough Moho interface is necessary to reduce precritical Pm</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/31336','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/31336"><span><span class="hlt">Field</span> computation of winds-aloft <span class="hlt">velocities</span> from single theodolite pilot balloon <span class="hlt">observations</span></span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Bill C. Ryan</p> <p>1976-01-01</p> <p>The ability to determine wind speeds and directions in the first few thousand meters of the atmosphere is important in many forestry operations such as smolce management, aircraft seeding and spraying, prescribed burning, and wildfire suppression. A hand-held electronic calculator can be used to compute winds aloft as balloon <span class="hlt">observations</span> are taken. Calculations can...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19760037453&hterms=insect+cells&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dinsect%2Bcells','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19760037453&hterms=insect+cells&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dinsect%2Bcells"><span>Vertical <span class="hlt">velocity</span> structure and geometry of clear air convective elements</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rowland, J. R.; Arnold, A.</p> <p>1975-01-01</p> <p>The paper discusses <span class="hlt">observations</span> of individual convective elements with a high-power narrow-beam scanning radar, an FM-CW radar, and an acoustic sounder, including the determination of the vertical air <span class="hlt">velocity</span> patterns of convective structures with the FM-CW radar and acoustic sounder. Data are presented which link the <span class="hlt">observed</span> <span class="hlt">velocity</span> structure and geometrical patterns to previously proposed models of boundary layer convection. It is shown that the high-power radar provides a clear three-dimensional picture of convective cells and <span class="hlt">fields</span> over a large area with a resolution of 150 m, where the convective cells are roughly spherical. Analysis of time-height records of the FM-CW radar and acoustic sounder confirms the downdraft-entrainment mechanism of the convective cell. The Doppler return of the acoustic sounder and the insect-trail slopes on FM-CW radar records are independent but redundant methods for obtaining the vertical <span class="hlt">velocity</span> patterns of convective structures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010PhPl...17a2505C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010PhPl...17a2505C"><span><span class="hlt">Observation</span> of energetic electron confinement in a largely stochastic reversed-<span class="hlt">field</span> pinch plasma</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Clayton, D. J.; Chapman, B. E.; O'Connell, R.; Almagri, A. F.; Burke, D. R.; Forest, C. B.; Goetz, J. A.; Kaufman, M. C.; Bonomo, F.; Franz, P.; Gobbin, M.; Piovesan, P.</p> <p>2010-01-01</p> <p>Runaway electrons with energies >100 keV are <span class="hlt">observed</span> with the appearance of an m =1 magnetic island in the core of otherwise stochastic Madison Symmetric Torus [Dexter et al., Fusion Technol. 19, 131 (1991)] reversed-<span class="hlt">field</span>-pinch plasmas. The island is associated with the innermost resonant tearing mode, which is usually the largest in the m =1 spectrum. The island appears over a range of mode spectra, from those with a weakly dominant mode to those, referred to as quasi single helicity, with a strongly dominant mode. In a stochastic <span class="hlt">field</span>, the rate of electron loss increases with electron parallel <span class="hlt">velocity</span>. Hence, high-energy electrons imply a region of reduced stochasticity. The global energy confinement time is about the same as in plasmas without high-energy electrons or an island in the core. Hence, the region of reduced stochasticity must be localized. Within a numerical reconstruction of the magnetic <span class="hlt">field</span> topology, high-energy electrons are substantially better confined inside the island, relative to the external region. Therefore, it is deduced that the island provides a region of reduced stochasticity and that the high-energy electrons are generated and well confined within this region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013APS..DFDA14005L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013APS..DFDA14005L"><span>A Comparison of 3D3C <span class="hlt">Velocity</span> Measurement Techniques</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>La Foy, Roderick; Vlachos, Pavlos</p> <p>2013-11-01</p> <p>The <span class="hlt">velocity</span> measurement fidelity of several 3D3C PIV measurement techniques including tomographic PIV, synthetic aperture PIV, plenoptic PIV, defocusing PIV, and 3D PTV are compared in simulations. A physically realistic ray-tracing algorithm is used to generate synthetic images of a standard calibration grid and of illuminated particle <span class="hlt">fields</span> advected by homogeneous isotropic turbulence. The simulated images for the tomographic, synthetic aperture, and plenoptic PIV cases are then used to create three-dimensional reconstructions upon which cross-correlations are performed to yield the measured <span class="hlt">velocity</span> <span class="hlt">field</span>. Particle tracking algorithms are applied to the images for the defocusing PIV and 3D PTV to directly yield the three-dimensional <span class="hlt">velocity</span> <span class="hlt">field</span>. In all cases the measured <span class="hlt">velocity</span> <span class="hlt">fields</span> are compared to one-another and to the true <span class="hlt">velocity</span> <span class="hlt">field</span> using several metrics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015IAUS..305..238M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015IAUS..305..238M"><span>Inhomogeneity and <span class="hlt">velocity</span> <span class="hlt">fields</span> effects on scattering polarization in solar prominences</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Milić, I.; Faurobert, M.</p> <p>2015-10-01</p> <p>One of the methods for diagnosing vector magnetic <span class="hlt">fields</span> in solar prominences is the so called "inversion" of <span class="hlt">observed</span> polarized spectral lines. This inversion usually assumes a fairly simple generative model and in this contribution we aim to study the possible systematic errors that are introduced by this assumption. On two-dimensional toy model of a prominence, we first demonstrate importance of multidimensional radiative transfer and horizontal inhomogeneities. These are able to induce a significant level of polarization in Stokes U, without the need for the magnetic <span class="hlt">field</span>. We then compute emergent Stokes spectrum from a prominence which is pervaded by the vector magnetic <span class="hlt">field</span> and use a simple, one-dimensional model to interpret these synthetic <span class="hlt">observations</span>. We find that inferred values for the magnetic <span class="hlt">field</span> vector generally differ from the original ones. Most importantly, the magnetic <span class="hlt">field</span> might seem more inclined than it really is.</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 gradients 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 <span class="hlt">field</span> 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 <span class="hlt">field</span> 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('http://adsabs.harvard.edu/abs/2014ASPC..481..287T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014ASPC..481..287T"><span>Low-Resolution Radial-<span class="hlt">Velocity</span> Monitoring of Pulsating sdBs in the Kepler <span class="hlt">Field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Telting, J.; Östensen, R.; Reed, M.; Kiæerad, F.; Farris, L.; Baran, A.; Oreiro, R.; O'Toole, S.</p> <p>2014-04-01</p> <p>We present preliminary results from an ongoing spectroscopic campaign to uncover the binary status of the 18 known pulsating subdwarf B stars and the one pulsating BHB star <span class="hlt">observed</span> with the Kepler spacecraft. During the 2010-2012 <span class="hlt">observing</span> seasons, we have used the KP4m Mayall, NOT, and WHT telescopes to obtain low-resolution (R˜2000-2500) Balmer-line spectroscopy of our sample stars. We applied a standard cross-correlation technique to derive radial <span class="hlt">velocities</span>, and find clear evidence for binarity in several of the pulsators, some of which were not previously known to be binaries.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22196044','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22196044"><span>A phenomenological retention tank model using settling <span class="hlt">velocity</span> distributions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Maruejouls, T; Vanrolleghem, P A; Pelletier, G; Lessard, P</p> <p>2012-12-15</p> <p>Many authors have <span class="hlt">observed</span> the influence of the settling <span class="hlt">velocity</span> distribution on the sedimentation process in retention tanks. However, the pollutants' behaviour in such tanks is not well characterized, especially with respect to their settling <span class="hlt">velocity</span> distribution. This paper presents a phenomenological modelling study dealing with the way by which the settling <span class="hlt">velocity</span> distribution of particles in combined sewage changes between entering and leaving an off-line retention tank. The work starts from a previously published model (Lessard and Beck, 1991) which is first implemented in a wastewater management modelling software, to be then tested with full-scale <span class="hlt">field</span> data for the first time. Next, its performance is improved by integrating the particle settling <span class="hlt">velocity</span> distribution and adding a description of the resuspension due to pumping for emptying the tank. Finally, the potential of the improved model is demonstrated by comparing the results for one more rain event. Copyright © 2011 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25215826','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25215826"><span>Statistical scaling of pore-scale Lagrangian <span class="hlt">velocities</span> in natural porous media.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Siena, M; Guadagnini, A; Riva, M; Bijeljic, B; Pereira Nunes, J P; Blunt, M J</p> <p>2014-08-01</p> <p>We investigate the scaling behavior of sample statistics of pore-scale Lagrangian <span class="hlt">velocities</span> in two different rock samples, Bentheimer sandstone and Estaillades limestone. The samples are imaged using x-ray computer tomography with micron-scale resolution. The scaling analysis relies on the study of the way qth-order sample structure functions (statistical moments of order q of absolute increments) of Lagrangian <span class="hlt">velocities</span> depend on separation distances, or lags, traveled along the mean flow direction. In the sandstone block, sample structure functions of all orders exhibit a power-law scaling within a clearly identifiable intermediate range of lags. Sample structure functions associated with the limestone block display two diverse power-law regimes, which we infer to be related to two overlapping spatially correlated structures. In both rocks and for all orders q, we <span class="hlt">observe</span> linear relationships between logarithmic structure functions of successive orders at all lags (a phenomenon that is typically known as extended power scaling, or extended self-similarity). The scaling behavior of Lagrangian <span class="hlt">velocities</span> is compared with the one exhibited by porosity and specific surface area, which constitute two key pore-scale geometric <span class="hlt">observables</span>. The statistical scaling of the local <span class="hlt">velocity</span> <span class="hlt">field</span> reflects the behavior of these geometric <span class="hlt">observables</span>, with the occurrence of power-law-scaling regimes within the same range of lags for sample structure functions of Lagrangian <span class="hlt">velocity</span>, porosity, and specific surface area.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20100026391&hterms=figueroa&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dfigueroa','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20100026391&hterms=figueroa&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dfigueroa"><span>Solar Wind Halo Formation by the Scattering of the Strahl via Direct Cluster/PEACE <span class="hlt">Observations</span> of the 3D <span class="hlt">Velocity</span> Distribution Function</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Figueroa-Vinas, Adolfo; Gurgiolo, Chris A.; Nieves-Chinchilla, Teresa; Goldstein, Melvyn L.</p> <p>2010-01-01</p> <p>It has been suggested by a number of authors that the solar wind electron halo can be formed by the scattering of the strahl. On frequent occasions we have <span class="hlt">observed</span> in electron angular skymaps (Phi/Theta-plots) of the electron 3D <span class="hlt">velocity</span> distribution functions) a bursty-filament of particles connecting the strahl to the solar wind core-halo. These are seen over a very limited energy range. When the magnetic <span class="hlt">field</span> is well off the nominal solar wind flow direction such filaments are inconsistent with any local forces and are probably the result of strong scattering. Furthermore, <span class="hlt">observations</span> indicates that the strahl component is frequently and significantly anisotropic (Tper/Tpal approx.2). This provides a possible free energy source for the excitation of whistler waves as a possible scattering mechanism. The empirical <span class="hlt">observational</span> evidence between the halo and the strahl suggests that the strahl population may be, at least in part, the source of the halo component.</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/2018MNRAS.473.3949B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.473.3949B"><span>Initial <span class="hlt">velocity</span> V-shapes of young asteroid families</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bolin, Bryce T.; Walsh, Kevin J.; Morbidelli, Alessandro; Delbó, Marco</p> <p>2018-01-01</p> <p>Ejection <span class="hlt">velocity</span> <span class="hlt">fields</span> of asteroid families are largely unconstrained due to the fact that members disperse relatively quickly on Myr time-scales by secular resonances and the Yarkovsky effect. The spreading of fragments in a by the Yarkovsky effect is indistinguishable from the spreading caused by the initial ejection of fragments. By examining families <20 Myr old, we can use the V-shape identification technique to separate family shapes that are due to the initial ejection <span class="hlt">velocity</span> <span class="hlt">field</span> and those that are due to the Yarkovsky effect. Asteroid families that are <20 Myr old provide an opportunity to study the <span class="hlt">velocity</span> <span class="hlt">field</span> of family fragments before they become too dispersed. Only the Karin family's initial <span class="hlt">velocity</span> <span class="hlt">field</span> has been determined and scales inversely with diameter, D-1. We have applied the V-shape identification technique to constrain young families' initial ejection <span class="hlt">velocity</span> <span class="hlt">fields</span> by measuring the curvature of their fragments' V-shape correlation in semimajor axis, a, versus D-1 space. Curvature from a straight line implies a deviation from a scaling of D-1. We measure the V-shape curvature of 11 young asteroid families including the 1993 FY12, Aeolia, Brangane, Brasilia, Clarissa, Iannini, Karin, Konig, Koronis(2), Theobalda and Veritas asteroid families. We find that the majority of asteroid families have initial ejection <span class="hlt">velocity</span> <span class="hlt">fields</span> consistent with ∼D-1 supporting laboratory impact experiments and computer simulations of disrupting asteroid parent bodies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24595169','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24595169"><span>Frictional <span class="hlt">velocity</span>-weakening in landslides on Earth and on other planetary bodies.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lucas, Antoine; Mangeney, Anne; Ampuero, Jean Paul</p> <p>2014-03-04</p> <p>One of the ultimate goals in landslide hazard assessment is to predict maximum landslide extension and <span class="hlt">velocity</span>. Despite much work, the physical processes governing energy dissipation during these natural granular flows remain uncertain. <span class="hlt">Field</span> <span class="hlt">observations</span> show that large landslides travel over unexpectedly long distances, suggesting low dissipation. Numerical simulations of landslides require a small friction coefficient to reproduce the extension of their deposits. Here, based on analytical and numerical solutions for granular flows constrained by remote-sensing <span class="hlt">observations</span>, we develop a consistent method to estimate the effective friction coefficient of landslides. This method uses a constant basal friction coefficient that reproduces the first-order landslide properties. We show that friction decreases with increasing volume or, more fundamentally, with increasing sliding <span class="hlt">velocity</span>. Inspired by frictional weakening mechanisms thought to operate during earthquakes, we propose an empirical <span class="hlt">velocity</span>-weakening friction law under a unifying phenomenological framework applicable to small and large landslides <span class="hlt">observed</span> on Earth and beyond.</p> </li> <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/2017AGUFM.C51B0976R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C51B0976R"><span><span class="hlt">Velocity</span> <span class="hlt">Field</span> of the McMurdo Shear Zone from Annual Three-Dimensional Ground Penetrating Radar Imaging and Crevasse Matching</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ray, L.; Jordan, M.; Arcone, S. A.; Kaluzienski, L. M.; Koons, P. O.; Lever, J.; Walker, B.; Hamilton, G. S.</p> <p>2017-12-01</p> <p>The McMurdo Shear Zone (MSZ) is a narrow, intensely crevassed strip tens of km long separating the Ross and McMurdo ice shelves (RIS and MIS) and an important pinning feature for the RIS. We derive local <span class="hlt">velocity</span> <span class="hlt">fields</span> within the MSZ from two consecutive annual ground penetrating radar (GPR) datasets that reveal complex firn and marine ice crevassing; no englacial features are evident. The datasets were acquired in 2014 and 2015 using robot-towed 400 MHz and 200 MHz GPR over a 5 km x 5.7 km grid. 100 west-to-east transects at 50 m spacing provide three-dimensional maps that reveal the length of many firn crevasses, and their year-to-year structural evolution. Hand labeling of crevasse cross sections near the MSZ western and eastern boundaries reveal matching firn and marine ice crevasses, and more complex and chaotic features between these boundaries. By matching crevasse features from year to year both on the eastern and western boundaries and within the chaotic region, marine ice crevasses along the western and eastern boundaries are shown to align directly with firn crevasses, and the local <span class="hlt">velocity</span> <span class="hlt">field</span> is estimated and compared with data from strain rate surveys and remote sensing. While remote sensing provides global <span class="hlt">velocity</span> <span class="hlt">fields</span>, crevasse matching indicates greater local complexity attributed to faulting, folding, and rotation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=magnetic+AND+particles&id=EJ1054981','ERIC'); return false;" href="https://eric.ed.gov/?q=magnetic+AND+particles&id=EJ1054981"><span>Revisiting the <span class="hlt">Velocity</span> Selector Problem with VPython</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>Milbourne, Jeff; Lim, Halson</p> <p>2015-01-01</p> <p>The <span class="hlt">velocity</span> selector is a classic first-year physics problem that demonstrates the influence of perpendicular electric and magnetic <span class="hlt">fields</span> on a charged particle. Traditionally textbooks introduce this problem in the context of balanced forces, often asking for <span class="hlt">field</span> strengths that would allow a charged particle, with a specific target <span class="hlt">velocity</span>,…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22520037-lamost-observations-kepler-field-spectral-classification-mkclass-code','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22520037-lamost-observations-kepler-field-spectral-classification-mkclass-code"><span>LAMOST <span class="hlt">OBSERVATIONS</span> IN THE KEPLER <span class="hlt">FIELD</span>: SPECTRAL CLASSIFICATION WITH THE MKCLASS CODE</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>Gray, R. O.; Corbally, C. J.; Cat, P. De</p> <p>2016-01-15</p> <p>The LAMOST-Kepler project was designed to obtain high-quality, low-resolution spectra of many of the stars in the Kepler <span class="hlt">field</span> with the Large Sky Area Multi Object Fiber Spectroscopic Telescope (LAMOST) spectroscopic telescope. To date 101,086 spectra of 80,447 objects over the entire Kepler <span class="hlt">field</span> have been acquired. Physical parameters, radial <span class="hlt">velocities</span>, and rotational <span class="hlt">velocities</span> of these stars will be reported in other papers. In this paper we present MK spectral classifications for these spectra determined with the automatic classification code MKCLASS. We discuss the quality and reliability of the spectral types and present histograms showing the frequency of the spectralmore » types in the main table organized according to luminosity class. Finally, as examples of the use of this spectral database, we compute the proportion of A-type stars that are Am stars, and identify 32 new barium dwarf candidates.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018TePhL..44..260P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018TePhL..44..260P"><span>Increasing Saturated Electron-Drift <span class="hlt">Velocity</span> in Donor-Acceptor Doped pHEMT Heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Protasov, D. Yu.; Gulyaev, D. V.; Bakarov, A. K.; Toropov, A. I.; Erofeev, E. V.; Zhuravlev, K. S.</p> <p>2018-03-01</p> <p><span class="hlt">Field</span> dependences of the electron-drift <span class="hlt">velocity</span> in typical pseudomorphic high-electron-mobility transistor (pHEMT) heteroepitaxial structures (HESs) and in those with donor-acceptor doped (DApHEMT) heterostructures with quantum-well (QW) depth increased by 0.8-0.9 eV with the aid of acceptor layers have been studied by a pulsed technique. It is established that the saturated electron-drift <span class="hlt">velocity</span> in DA-pHEMT-HESs is 1.2-1.3 times greater than that in the usual pHEMT-HESs. The electroluminescence (EL) spectra of DA-pHEMT-HESs do not contain emission bands related to the recombination in widebandgap layers (QW barriers). The EL intensity in these HESs is not saturated with increasing electric <span class="hlt">field</span>. This is indicative of a suppressed real-space transfer of hot electrons from QW to barrier layers, which accounts for the <span class="hlt">observed</span> increase in the saturated electron-drift <span class="hlt">velocity</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22676050-compactly-supported-linearised-observables-single-field-inflation','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22676050-compactly-supported-linearised-observables-single-field-inflation"><span>Compactly supported linearised <span class="hlt">observables</span> in single-<span class="hlt">field</span> inflation</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>Fröob, Markus B.; Higuchi, Atsushi; Hack, Thomas-Paul, E-mail: mbf503@york.ac.uk, E-mail: thomas-paul.hack@itp.uni-leipzig.de, E-mail: atsushi.higuchi@york.ac.uk</p> <p></p> <p>We investigate the gauge-invariant <span class="hlt">observables</span> constructed by smearing the graviton and inflaton <span class="hlt">fields</span> by compactly supported tensors at linear order in general single-<span class="hlt">field</span> inflation. These <span class="hlt">observables</span> correspond to gauge-invariant quantities that can be measured locally. In particular, we show that these <span class="hlt">observables</span> are equivalent to (smeared) local gauge-invariant <span class="hlt">observables</span> such as the linearised Weyl tensor, which have better infrared properties than the graviton and inflaton <span class="hlt">fields</span>. Special cases include the equivalence between the compactly supported gauge-invariant graviton <span class="hlt">observable</span> and the smeared linearised Weyl tensor in Minkowski and de Sitter spaces. Our results indicate that the infrared divergences in the tensormore » and scalar perturbations in single-<span class="hlt">field</span> inflation have the same status as in de Sitter space and are both a gauge artefact, in a certain technical sense, at tree level.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70042784','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70042784"><span>Using cluster analysis to organize and explore regional GPS <span class="hlt">velocities</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Simpson, Robert W.; Thatcher, Wayne; Savage, James C.</p> <p>2012-01-01</p> <p>Cluster analysis offers a simple visual exploratory tool for the initial investigation of regional Global Positioning System (GPS) <span class="hlt">velocity</span> <span class="hlt">observations</span>, which are providing increasingly precise mappings of actively deforming continental lithosphere. The deformation <span class="hlt">fields</span> from dense regional GPS networks can often be concisely described in terms of relatively coherent blocks bounded by active faults, although the choice of blocks, their number and size, can be subjective and is often guided by the distribution of known faults. To illustrate our method, we apply cluster analysis to GPS <span class="hlt">velocities</span> from the San Francisco Bay Region, California, to search for spatially coherent patterns of deformation, including evidence of block-like behavior. The clustering process identifies four robust groupings of <span class="hlt">velocities</span> that we identify with four crustal blocks. Although the analysis uses no prior geologic information other than the GPS <span class="hlt">velocities</span>, the cluster/block boundaries track three major faults, both locked and creeping.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhPl...25c2104H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhPl...25c2104H"><span><span class="hlt">Observations</span> of a <span class="hlt">field</span>-aligned ion/ion-beam instability in a magnetized laboratory plasma</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Heuer, P. V.; Weidl, M. S.; Dorst, R. S.; Schaeffer, D. B.; Bondarenko, A. S.; Tripathi, S. K. P.; Van Compernolle, B.; Vincena, S.; Constantin, C. G.; Niemann, C.; Winske, D.</p> <p>2018-03-01</p> <p>Collisionless coupling between super Alfvénic ions and an ambient plasma parallel to a background magnetic <span class="hlt">field</span> is mediated by a set of electromagnetic ion/ion-beam instabilities including the resonant right hand instability (RHI). To study this coupling and its role in parallel shock formation, a new experimental configuration at the University of California, Los Angeles utilizes high-energy and high-repetition-rate lasers to create a super-Alfvénic <span class="hlt">field</span>-aligned debris plasma within an ambient plasma in the Large Plasma Device. We used a time-resolved fluorescence monochromator and an array of Langmuir probes to characterize the laser plasma <span class="hlt">velocity</span> distribution and density. The debris ions were <span class="hlt">observed</span> to be sufficiently super-Alfvénic and dense to excite the RHI. Measurements with magnetic flux probes exhibited a right-hand circularly polarized frequency chirp consistent with the excitation of the RHI near the laser target. We compared measurements to 2D hybrid simulations of the experiment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013WRR....49.3093C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013WRR....49.3093C"><span>River <span class="hlt">velocities</span> from sequential multispectral remote sensing images</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Wei; Mied, Richard P.</p> <p>2013-06-01</p> <p>We address the problem of extracting surface <span class="hlt">velocities</span> from a pair of multispectral remote sensing images over rivers using a new nonlinear multiple-tracer form of the global optimal solution (GOS). The derived <span class="hlt">velocity</span> <span class="hlt">field</span> is a valid solution across the image domain to the nonlinear system of equations obtained by minimizing a cost function inferred from the conservation constraint equations for multiple tracers. This is done by deriving an iteration equation for the <span class="hlt">velocity</span>, based on the multiple-tracer displaced frame difference equations, and a local approximation to the <span class="hlt">velocity</span> <span class="hlt">field</span>. The number of <span class="hlt">velocity</span> equations is greater than the number of <span class="hlt">velocity</span> components, and thus overly constrain the solution. The iterative technique uses Gauss-Newton and Levenberg-Marquardt methods and our own algorithm of the progressive relaxation of the over-constraint. We demonstrate the nonlinear multiple-tracer GOS technique with sequential multispectral Landsat and ASTER images over a portion of the Potomac River in MD/VA, and derive a dense <span class="hlt">field</span> of accurate <span class="hlt">velocity</span> vectors. We compare the GOS river <span class="hlt">velocities</span> with those from over 12 years of data at four NOAA reference stations, and find good agreement. We discuss how to find the appropriate spatial and temporal resolutions to allow optimization of the technique for specific rivers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AAS...22915504R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AAS...22915504R"><span>Quality Control of The Miniature Exoplanet Radio <span class="hlt">Velocity</span> Array(MINERVA)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rivera García, Kevin O.; Eastman, Jason D.</p> <p>2017-01-01</p> <p>The MINiature Exoplanet Radial <span class="hlt">Velocity</span> Array, also known as MINERVA , is a network of four robotic 0.7 meter telescopes that is conducting a Radial <span class="hlt">Velocity</span> survey of the nearest, brightest stars in search of small and rocky exoplanets. The robotic telescope array is located in Fred Lawrence Whipple Observatory in Arizona. MINERVA began science operations in 2015 and we are constantly improving its <span class="hlt">observing</span> efficiency. We will describe performance statistics that we have developed in Python to proactively identify problems before they impede <span class="hlt">observations</span>. We have written code to monitor the pointing error for each telescope to ensure it will always be able to acquire a target in the 3 arcminute <span class="hlt">field</span> of view of its acquisition camera, but there are still some issues that need to be identified. The end goal for this research is to automatically address any common malfunction that may cause the <span class="hlt">observation</span> to fail and ultimately improve our <span class="hlt">observing</span> efficiency.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007AIPC..955..919H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AIPC..955..919H"><span>Low <span class="hlt">Velocity</span> Detonation of Nitromethane Affected by Precursor Shock Waves Propagating in Various Container Materials</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hamashima, H.; Osada, A.; Itoh, S.; Kato, Y.</p> <p>2007-12-01</p> <p>It is well known that some liquid explosives have two detonation behaviors, high <span class="hlt">velocity</span> detonation (HVD) or low <span class="hlt">velocity</span> detonation (LVD) can propagate. A physical model to describe the propagation mechanism of LVD in liquid explosives was proposed that LVD is not a self-reactive detonation, but rather a supported-reactive detonation from the cavitation <span class="hlt">field</span> generated by precursor shock waves. However, the detailed structure of LVD in liquid explosives has not yet been clarified. In this study, high-speed photography was used to investigate the effects of the precursor shock waves propagating in various container materials for LVD in nitromethane (NM). Stable LVD was not <span class="hlt">observed</span> in all containers, although transient LVD was <span class="hlt">observed</span>. A very complicated structure of LVD was <span class="hlt">observed</span>: the interaction of multiple precursor shock waves, multiple oblique shock waves, and the cavitation <span class="hlt">field</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007APS..SHK.E1007H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007APS..SHK.E1007H"><span>Low <span class="hlt">Velocity</span> Detonation of Nitromethane Affected by Precursor Shock Waves Propagating in Various Container Materials</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hamashima, Hideki; Osada, Akinori; Kato, Yukio; Itoh, Shigeru</p> <p>2007-06-01</p> <p>It is well known that some liquid explosives have two detonation behaviors, high <span class="hlt">velocity</span> detonation (HVD) or low <span class="hlt">velocity</span> detonation (LVD) can propagate. A physical model to describe the propagation mechanism of LVD in liquid explosives was proposed that LVD is not a self-reactive detonation, but rather a supported-reactive detonation from the cavitation <span class="hlt">field</span> generated by precursor shock waves. However, the detailed structure of LVD in liquid explosives has not yet been clarified. In this study, high-speed photography was used to investigate the effects of the precursor shock waves propagating in various container materials for LVD in nitromethane (NM). Stable LVD was not <span class="hlt">observed</span> in all containers, although transient LVD was <span class="hlt">observed</span>. A very complicated structure of LVD was <span class="hlt">observed</span>: the interaction of multiple precursor shock waves, multiple oblique shock waves, and the cavitation <span class="hlt">field</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JSeis..20.1207F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JSeis..20.1207F"><span>Cohesive zone length of metagabbro at supershear 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>Fukuyama, Eiichi; Xu, Shiqing; Yamashita, Futoshi; Mizoguchi, Kazuo</p> <p>2016-10-01</p> <p>We investigated the shear strain <span class="hlt">field</span> ahead of a supershear rupture. The strain array data along the sliding fault surfaces were obtained during the large-scale biaxial friction experiments at the National Research Institute for Earth Science and Disaster Resilience. These friction experiments were done using a pair of meter-scale metagabbro rock specimens whose simulated fault area was 1.5 m × 0.1 m. A 2.6-MPa normal stress was applied with loading <span class="hlt">velocity</span> of 0.1 mm/s. Near-fault strain was measured by 32 two-component semiconductor strain gauges installed at an interval of 50 mm and 10 mm off the fault and recorded at an interval of 1 MHz. Many stick-slip events were <span class="hlt">observed</span> in the experiments. We chose ten unilateral rupture events that propagated with supershear rupture <span class="hlt">velocity</span> without preceding foreshocks. Focusing on the rupture front, stress concentration was <span class="hlt">observed</span> and sharp stress drop occurred immediately inside the ruptured area. The temporal variation of strain array data is converted to the spatial variation of strain assuming a constant rupture <span class="hlt">velocity</span>. We picked up the peak strain and zero-crossing strain locations to measure the cohesive zone length. By compiling the stick-slip event data, the cohesive zone length is about 50 mm although it scattered among the events. We could not see any systematic variation at the location but some dependence on the rupture <span class="hlt">velocity</span>. The cohesive zone length decreases as the rupture <span class="hlt">velocity</span> increases, especially larger than √{2} times the shear wave <span class="hlt">velocity</span>. This feature is consistent with the theoretical prediction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017A%26A...601A.103A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017A%26A...601A.103A"><span>Differences between Doppler <span class="hlt">velocities</span> of ions and neutral atoms in a solar prominence</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Anan, T.; Ichimoto, K.; Hillier, A.</p> <p>2017-05-01</p> <p>Context. In astrophysical systems with partially ionized plasma, the motion of ions is governed by the magnetic <span class="hlt">field</span> while the neutral particles can only feel the magnetic <span class="hlt">field</span>'s Lorentz force indirectly through collisions with ions. The drift in the <span class="hlt">velocity</span> between ionized and neutral species plays a key role in modifying important physical processes such as magnetic reconnection, damping of magnetohydrodynamic waves, transport of angular momentum in plasma through the magnetic <span class="hlt">field</span>, and heating. Aims: This paper aims to investigate the differences between Doppler <span class="hlt">velocities</span> of calcium ions and neutral hydrogen in a solar prominence to look for <span class="hlt">velocity</span> differences between the neutral and ionized species. Methods: We simultaneously <span class="hlt">observed</span> spectra of a prominence over an active region in H I 397 nm, H I 434 nm, Ca II 397 nm, and Ca II 854 nm using a high dispersion spectrograph of the Domeless Solar Telescope at Hida observatory. We compared the Doppler <span class="hlt">velocities</span>, derived from the shift of the peak of the spectral lines presumably emitted from optically-thin plasma. Results: There are instances when the difference in <span class="hlt">velocities</span> between neutral atoms and ions is significant, for example 1433 events ( 3% of sets of compared profiles) with a difference in <span class="hlt">velocity</span> between neutral hydrogen atoms and calcium ions greater than 3σ of the measurement error. However, we also found significant differences between the Doppler <span class="hlt">velocities</span> of two spectral lines emitted from the same species, and the probability density functions of <span class="hlt">velocity</span> difference between the same species is not significantly different from those between neutral atoms and ions. Conclusions: We interpreted the difference of Doppler <span class="hlt">velocities</span> as being a result of the motions of different components in the prominence along the line of sight, rather than the decoupling of neutral atoms from plasma. The movie attached to Fig. 1 is available at http://www.aanda.org</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016cosp...41E1749S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016cosp...41E1749S"><span>Prominence plasma and magnetic <span class="hlt">field</span> structure - A coordinated <span class="hlt">observation</span> with IRIS, Hinode and THEMIS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schmieder, Brigitte; Labrosse, Nicolas; Levens, Peter; Lopez Ariste, Arturo</p> <p>2016-07-01</p> <p>During an international campaign in 2014, utilising both space-based (IRIS and Hinode) and ground-based (THEMIS) instruments, we focused on <span class="hlt">observing</span> prominences. We compare IRIS <span class="hlt">observations</span> with those of Hinode (EIS and SOT) in order to build a more complete picture of the prominence structure for a quiescent prominence <span class="hlt">observed</span> on 15 July 2014, identified to have tornado-like structure. THEMIS provides valuable information on the orientation and strength of the internal magnetic <span class="hlt">field</span>. Here we find there is almost ubiquitously horizontal <span class="hlt">field</span> with respect to the local limb, with possibly a turbulent component. The Mg II lines form the majority of our IRIS analysis, with a mixture of reversed and non-reversed profiles present in the prominence spectra. Comparing the differences between the Mg II data from IRIS and the Ca II images from Hinode/SOT provides an intriguing insight into the prominence legs in these channels. We present plasma diagnostics from IRIS, with line of sight <span class="hlt">velocities</span> of around 10 km/s in either direction along the magnetic loops of material in the front of the prominence, and line widths comparable to those found for prominences by previous authors (e.g. Schmieder et al. 2014). We also take a look into the lines formed at higher, coronal plasma temperatures, as seen by Hinode/EIS, to compare plasma structures at a full range of temperatures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22661379-synthetic-observations-magnetic-fields-protostellar-cores','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22661379-synthetic-observations-magnetic-fields-protostellar-cores"><span>SYNTHETIC <span class="hlt">OBSERVATIONS</span> OF MAGNETIC <span class="hlt">FIELDS</span> IN PROTOSTELLAR CORES</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>Lee, Joyce W. Y.; Hull, Charles L. H.; Offner, Stella S. R., E-mail: chat.hull@cfa.harvard.edu, E-mail: jwyl1g12@soton.ac.uk</p> <p></p> <p>The role of magnetic <span class="hlt">fields</span> in the early stages of star formation is not well constrained. In order to discriminate between different star formation models, we analyze 3D magnetohydrodynamic simulations of low-mass cores and explore the correlation between magnetic <span class="hlt">field</span> orientation and outflow orientation over time. We produce synthetic <span class="hlt">observations</span> of dust polarization at resolutions comparable to millimeter-wave dust polarization maps <span class="hlt">observed</span> by the Combined Array for Research in Millimeter-wave Astronomy and compare these with 2D visualizations of projected magnetic <span class="hlt">field</span> and column density. Cumulative distribution functions of the projected angle between the magnetic <span class="hlt">field</span> and outflow show different degreesmore » of alignment in simulations with differing mass-to-flux ratios. The distribution function for the less magnetized core agrees with <span class="hlt">observations</span> finding random alignment between outflow and <span class="hlt">field</span> orientations, while the more magnetized core exhibits stronger alignment. We find that fractional polarization increases when the system is viewed such that the magnetic <span class="hlt">field</span> is close to the plane of the sky, and the values of fractional polarization are consistent with <span class="hlt">observational</span> measurements. The simulation outflow, which reflects the underlying angular momentum of the accreted gas, changes direction significantly over over the first ∼0.1 Myr of evolution. This movement could lead to the <span class="hlt">observed</span> random alignment between outflows and the magnetic <span class="hlt">fields</span> in protostellar cores.« less</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('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 <span class="hlt">field</span> of view (˜11'×11'), SITELLE is the only integral <span class="hlt">field</span> 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 gradient 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> </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('https://ntrs.nasa.gov/search.jsp?R=19820028656&hterms=deutsche+forschungsgemeinschaft&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Ddeutsche%2Bforschungsgemeinschaft','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19820028656&hterms=deutsche+forschungsgemeinschaft&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Ddeutsche%2Bforschungsgemeinschaft"><span>On the <span class="hlt">velocity</span> distribution of ion jets during substorm recovery</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Birn, J.; Forbes, T. G.; Hones, E. W., Jr.; Bame, S. J.; Paschmann, G.</p> <p>1981-01-01</p> <p>The <span class="hlt">velocity</span> distribution of earthward jetting ions that are <span class="hlt">observed</span> principally during substorm recovery by satellites at approximately 15-35 earth radii in the magnetotail is quantitatively compared with two different theoretical models - the 'adiabatic deformation' of an initially flowing Maxwellian moving into higher magnetic <span class="hlt">field</span> strength (model A) and the <span class="hlt">field</span>-aligned electrostatic acceleration of an initially nonflowing isotropic Maxwellian including adiabatic deformation effects (model B). The assumption is made that the ions are protons or, more generally, that they consist of only one species. It is found that both models can explain the often <span class="hlt">observed</span> concave-convex shape of isodensity contours of the distribution function.</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('https://eric.ed.gov/?q=Are+AND+original+AND+AJ+AND+Quartmain+AND++AND+Michaels+AND+real+AND+Bio+AND+dad+AND+k&id=EJ1168703','ERIC'); return false;" href="https://eric.ed.gov/?q=Are+AND+original+AND+AJ+AND+Quartmain+AND++AND+Michaels+AND+real+AND+Bio+AND+dad+AND+k&id=EJ1168703"><span>Measuring Average Angular <span class="hlt">Velocity</span> with a Smartphone Magnetic <span class="hlt">Field</span> Sensor</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>Pili, Unofre; Violanda, Renante</p> <p>2018-01-01</p> <p>The angular <span class="hlt">velocity</span> of a spinning object is, by standard, measured using a device called a tachometer. However, by directly using it in a classroom setting, the activity is likely to appear as less instructive and less engaging. Indeed, some alternative classroom-suitable methods for measuring angular <span class="hlt">velocity</span> have been presented. In this paper,…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.4094N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.4094N"><span>Correlation between atmospheric electric <span class="hlt">fields</span> and cloud cover using a <span class="hlt">field</span> mill and cloud <span class="hlt">observation</span> data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nakamori, Kota; Suzuki, Yasuki; Ohya, Hiroyo; Takano, Toshiaki; Kawamura, Yohei; Nakata, Hiroyuki; Yamashita, Kozo</p> <p>2017-04-01</p> <p>It is known that lightning and precipitations of rain droplets generated from thunderclouds are a generator of global atmospheric electric circuit. In the fair weather, the atmospheric electric <span class="hlt">fields</span> (AEF) are downward (positive), while they are upward (negative) during lightning and precipitations. However, the correlations between the AEF, and the cloud parameters such as cloud cover, weather phenomenon, have been not revealed quantitatively yet. In this study, we investigate the correlations between the AEF and the cloud parameters, weather phenomenon using a <span class="hlt">field</span> mill, the 95 GHz-FALCON (FMCW Radar for Cloud <span class="hlt">Observations</span>)-I and all-sky camera <span class="hlt">observations</span>. In this study, we installed a Boltek <span class="hlt">field</span> mill on the roof of our building in Chiba University, Japan, (Geographic coordinate: 35.63 degree N, 140.10 degree E, the sea level: 55 m) on the first June, 2016. The sampling time of the AEF is 0.5 s. On the other hand, the FALCON-I has <span class="hlt">observed</span> the cloud parameters far from about 76 m of the <span class="hlt">field</span> mill throughout 24 hours every day. The vertical cloud profiles and the Doppler <span class="hlt">velocity</span> of cloud particles can be derived by the FALCON-I with high distance resolutions (48.8 m) (Takano et al., 2010). In addition, the images of the clouds and precipitations are recorded with 30-s sampling by an all-sky camera using a CCD camera on the same roof during 05:00-22:00 LT every day. The distance between the <span class="hlt">field</span> mill and the all-sky camera is 3.75 m. During 08:30 UT - 10:30 UT, on 4 July, 2016, we found the variation of the AEF due to the approach of thundercloud. The variation consisted of two patterns. One was slow variation due to the movement of thunderclouds, and the other was rapid variation associated with lightning discharges. As for the movement of thunderclouds, the AEF increased when the anvil was located over the <span class="hlt">field</span> mill, which was opposite direction of the previous studies. This change might be due to the positive charges in the upper anvil more than 14 km</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1932c0025M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1932c0025M"><span>Low <span class="hlt">velocity</span> impact of 6082-T6 aluminum plates</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mocian, Oana Alexandra; Constantinescu, Dan Mihai; Sandu, Marin; Sorohan, Ştefan</p> <p>2018-02-01</p> <p>The low <span class="hlt">velocity</span> domain covers vehicle impacts, ship collisions and even accidentally tool drops. Even though more and more research is needed into these <span class="hlt">fields</span>, most of the papers concerning impact problems focus on impact at medium and high <span class="hlt">velocities</span>. Understanding the behavior of structures subjected to low <span class="hlt">velocity</span> impact is of major importance when referring to impact resistance and damage tolerance. The paper presents an experimental and numerical investigation on the low <span class="hlt">velocity</span> behavior of 6082-T6 aluminum plates. Impact tests were performed using an Instron Ceast 9340 drop-weight testing machine. In the experimental procedure, square plates were mounted on a circular support, fixed with a pneumatic clamping system and impacted with a hemispherical steel projectile. Specimens were impacted at constant weight and different impact <span class="hlt">velocities</span>. The effect of different impact energies was investigated. The impact event was then simulated using the nonlinear finite element code LS_DYNA in order to determine the effect of strain rate upon the mechanical behavior of the aluminum plates. Moreover, in order to capture the exact behavior of the material, a special attention has been given to the selection of the correct material model and its parameters, which, in large extent, depend on the <span class="hlt">observed</span> behavior of the aluminum plate during the test and the actual response of the plate under simulation. The numerical predictions are compared with the experimental <span class="hlt">observations</span> and the applicability of the numerical model for further researches is analyzed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005AtmRe..74...89M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005AtmRe..74...89M"><span>Measured and modeled dry deposition <span class="hlt">velocities</span> over the ESCOMPTE area</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Michou, M.; Laville, P.; Serça, D.; Fotiadi, A.; Bouchou, P.; Peuch, V.-H.</p> <p>2005-03-01</p> <p>Measurements of the dry deposition <span class="hlt">velocity</span> of ozone have been made by the eddy correlation method during ESCOMPTE (Etude sur Site pour COntraindre les Modèles de Pollution atmosphérique et de Transport d'Emissions). The strong local variability of natural ecosystems was sampled over several weeks in May, June and July 2001 for four sites with varying surface characteristics. The sites included a maize <span class="hlt">field</span>, a Mediterranean forest, a Mediterranean shrub-land, and an almost bare soil. Measurements of nitrogen oxide deposition fluxes by the relaxed eddy correlation method have also been carried out at the same bare soil site. An evaluation of the deposition <span class="hlt">velocities</span> computed by the surface module of the multi-scale Chemistry and Transport Model MOCAGE is presented. This module relies on a resistance approach, with a detailed treatment of the stomatal contribution to the surface resistance. Simulations at the finest model horizontal resolution (around 10 km) are compared to <span class="hlt">observations</span>. If the seasonal variations are in agreement with the literature, comparisons between raw model outputs and <span class="hlt">observations</span>, at the different measurement sites and for the specific <span class="hlt">observing</span> periods, are contrasted. As the simulated meteorology at the scale of 10 km nicely captures the <span class="hlt">observed</span> situations, the default set of surface characteristics (averaged at the resolution of a grid cell) appears to be one of the main reasons for the discrepancies found with <span class="hlt">observations</span>. For each case, sensitivity studies have been performed in order to see the impact of adjusting the surface characteristics to the <span class="hlt">observed</span> ones, when available. Generally, a correct agreement with the <span class="hlt">observations</span> of deposition <span class="hlt">velocities</span> is obtained. This advocates for a sub-grid scale representation of surface characteristics for the simulation of dry deposition <span class="hlt">velocities</span> over such a complex area. Two other aspects appear in the discussion. Firstly, the strong influence of the soil water content to the plant</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRA..122.8014B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..122.8014B"><span>Ion <span class="hlt">velocity</span> distributions in dipolarization events: Distributions in the central plasma sheet</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Birn, J.; Runov, A.; Zhou, X.-Z.</p> <p>2017-08-01</p> <p>Using combined MHD/test particle simulations, we further explore characteristic ion <span class="hlt">velocity</span> distributions in the central plasma sheet (CPS) in relation to dipolarization events. Distributions in the CPS within the dipolarized flux bundle (DFB) that follows the passage of a dipolarization front typically show two opposing low subthermal-energy beams with a ring-like component perpendicular to the magnetic <span class="hlt">field</span> at about twice the thermal energy. The dominance of the perpendicular anisotropy and a <span class="hlt">field</span>-aligned peak at lower energy agree qualitatively with ion distribution functions derived from "Time History of Events and Macroscale Interactions during Substorms" <span class="hlt">observations</span>. At locations somewhat off the equatorial plane the <span class="hlt">field</span>-aligned peaks are shifted by a <span class="hlt">field</span>-aligned component of the bulk flow, such that one peak becomes centered near zero net <span class="hlt">velocity</span>, which makes it less likely to be <span class="hlt">observed</span>. The origins of the <span class="hlt">field</span>-aligned peaks are low-energy lobe (or near plasma sheet boundary layer) regions, while the ring distribution originates mostly from thermal plasma sheet particles on extended <span class="hlt">field</span> lines. The acceleration mechanisms are also quite different: the beam ions are accelerated first by the E × B drift motion of the DFB and then by a slingshot effect of the earthward convecting DFB (akin to first-order Fermi, type B, acceleration), which causes an increase in <span class="hlt">field</span>-aligned speed. In contrast, the ring particles are accelerated by successive, betatron-like acceleration after entering the high electric <span class="hlt">field</span> region of an earthward propagating DFB.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008A%26A...481..367H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008A%26A...481..367H"><span>Dissipative structures of diffuse molecular gas. III. Small-scale intermittency of intense <span class="hlt">velocity</span>-shears</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hily-Blant, P.; Falgarone, E.; Pety, J.</p> <p>2008-04-01</p> <p>Aims: We further characterize the structures tentatively identified on thermal and chemical grounds as the sites of dissipation of turbulence in molecular clouds (Papers I and II). Methods: Our study is based on two-point statistics of line centroid <span class="hlt">velocities</span> (CV), computed from three large 12CO maps of two <span class="hlt">fields</span>. We build the probability density functions (PDF) of the CO line centroid <span class="hlt">velocity</span> increments (CVI) over lags varying by an order of magnitude. Structure functions of the line CV are computed up to the 6th order. We compare these statistical properties in two translucent parsec-scale <span class="hlt">fields</span> embedded in different large-scale environments, one far from virial balance and the other virialized. We also address their scale dependence in the former, more turbulent, <span class="hlt">field</span>. Results: The statistical properties of the line CV bear the three signatures of intermittency in a turbulent <span class="hlt">velocity</span> <span class="hlt">field</span>: (1) the non-Gaussian tails in the CVI PDF grow as the lag decreases, (2) the departure from Kolmogorov scaling of the high-order structure functions is more pronounced in the more turbulent <span class="hlt">field</span>, (3) the positions contributing to the CVI PDF tails delineate narrow filamentary structures (thickness ~0.02 pc), uncorrelated to dense gas structures and spatially coherent with thicker ones (~0.18 pc) <span class="hlt">observed</span> on larger scales. We show that the largest CVI trace sharp variations of the extreme CO linewings and that they actually capture properties of the underlying <span class="hlt">velocity</span> <span class="hlt">field</span>, uncontaminated by density fluctuations. The confrontation with theoretical predictions leads us to identify these small-scale filamentary structures with extrema of <span class="hlt">velocity</span>-shears. We estimate that viscous dissipation at the 0.02 pc-scale in these structures is up to 10 times higher than average, consistent with their being associated with gas warmer than the bulk. Last, their average direction is parallel (or close) to that of the local magnetic <span class="hlt">field</span> projection. Conclusions: Turbulence in these</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21693999','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21693999"><span>Microscopic theory of longitudinal sound <span class="hlt">velocity</span> in charge ordered manganites.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rout, G C; Panda, S</p> <p>2009-10-14</p> <p>A microscopic theory of longitudinal sound <span class="hlt">velocity</span> in a manganite system is reported here. The manganite system is described by a model Hamiltonian consisting of charge density wave (CDW) interaction in the e(g) band, an exchange interaction between spins of the itinerant e(g) band electrons and the core t(2g) electrons, and the Heisenberg interaction of the core level spins. The magnetization and the CDW order parameters are considered within mean-<span class="hlt">field</span> approximations. The phonon Green's function was calculated by Zubarev's technique and hence the longitudinal <span class="hlt">velocity</span> of sound was finally calculated for the manganite system. The results show that the elastic spring involved in the <span class="hlt">velocity</span> of sound exhibits strong stiffening in the CDW phase with a decrease in temperature as <span class="hlt">observed</span> in experiments.</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/2012AAS...21944002M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AAS...21944002M"><span>Measuring the Power Spectrum with Peculiar <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>Macaulay, Edward; Feldman, H. A.; Ferreira, P. G.; Jaffe, A. H.; Agarwal, S.; Hudson, M. J.; Watkins, R.</p> <p>2012-01-01</p> <p>The peculiar <span class="hlt">velocities</span> of galaxies are an inherently valuable cosmological probe, providing an unbiased estimate of the distribution of matter on scales much larger than the depth of the survey. Much research interest has been motivated by the high dipole moment of our local peculiar <span class="hlt">velocity</span> <span class="hlt">field</span>, which suggests a large scale excess in the matter power spectrum, and can appear to be in some tension with the LCDM model. We use a composite catalogue of 4,537 peculiar <span class="hlt">velocity</span> measurements with a characteristic depth of 33 h-1 Mpc to estimate the matter power spectrum. We compare the constraints with this method, directly studying the full peculiar <span class="hlt">velocity</span> catalogue, to results from Macaulay et al. (2011), studying minimum variance moments of the <span class="hlt">velocity</span> <span class="hlt">field</span>, as calculated by Watkins, Feldman & Hudson (2009) and Feldman, Watkins & Hudson (2010). We find good agreement with the LCDM model on scales of k > 0.01 h Mpc-1. We find an excess of power on scales of k < 0.01 h Mpc-1, although with a 1 sigma uncertainty which includes the LCDM model. We find that the uncertainty in the excess at these scales is larger than an alternative result studying only moments of the <span class="hlt">velocity</span> <span class="hlt">field</span>, which is due to the minimum variance weights used to calculate the moments. At small scales, we are able to clearly discriminate between linear and nonlinear clustering in simulated peculiar <span class="hlt">velocity</span> catalogues, and find some evidence (although less clear) for linear clustering in the real peculiar <span class="hlt">velocity</span> data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012MNRAS.425.1709M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012MNRAS.425.1709M"><span>Power spectrum estimation from peculiar <span class="hlt">velocity</span> catalogues</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Macaulay, E.; Feldman, H. A.; Ferreira, P. G.; Jaffe, A. H.; Agarwal, S.; Hudson, M. J.; Watkins, R.</p> <p>2012-09-01</p> <p>The peculiar <span class="hlt">velocities</span> of galaxies are an inherently valuable cosmological probe, providing an unbiased estimate of the distribution of matter on scales much larger than the depth of the survey. Much research interest has been motivated by the high dipole moment of our local peculiar <span class="hlt">velocity</span> <span class="hlt">field</span>, which suggests a large-scale excess in the matter power spectrum and can appear to be in some tension with the Λ cold dark matter (ΛCDM) model. We use a composite catalogue of 4537 peculiar <span class="hlt">velocity</span> measurements with a characteristic depth of 33 h-1 Mpc to estimate the matter power spectrum. We compare the constraints with this method, directly studying the full peculiar <span class="hlt">velocity</span> catalogue, to results by Macaulay et al., studying minimum variance moments of the <span class="hlt">velocity</span> <span class="hlt">field</span>, as calculated by Feldman, Watkins & Hudson. We find good agreement with the ΛCDM model on scales of k > 0.01 h Mpc-1. We find an excess of power on scales of k < 0.01 h Mpc-1 with a 1σ uncertainty which includes the ΛCDM model. We find that the uncertainty in excess at these scales is larger than an alternative result studying only moments of the <span class="hlt">velocity</span> <span class="hlt">field</span>, which is due to the minimum variance weights used to calculate the moments. At small scales, we are able to clearly discriminate between linear and non-linear clustering in simulated peculiar <span class="hlt">velocity</span> catalogues and find some evidence (although less clear) for linear clustering in the real peculiar <span class="hlt">velocity</span> data.</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 <span class="hlt">field</span> 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 gradient (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 <span class="hlt">field</span> 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 <span class="hlt">field</span> caused a small noticeable impact on the growth rate of electrostatic electron cyclotron instability. Density gradient 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/2011Tectp.503...34K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011Tectp.503...34K"><span>Stress anisotropy and <span class="hlt">velocity</span> anisotropy in low porosity shale</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kuila, U.; Dewhurst, D. N.; Siggins, A. F.; Raven, M. D.</p> <p>2011-04-01</p> <p>Shales are known for often marked intrinsic anisotropy of many of their properties, including strength, permeability and <span class="hlt">velocity</span> for example. In addition, it is well known that anisotropic stress <span class="hlt">fields</span> can also have a significant impact on anisotropy of <span class="hlt">velocity</span>, even in an isotropic medium. This paper sets out to investigate the ultrasonic <span class="hlt">velocity</span> response of well-characterised low porosity shales from the Officer Basin in Western Australia to both isotropic and anisotropic stress <span class="hlt">fields</span> and to evaluate the <span class="hlt">velocity</span> response to the changing stress <span class="hlt">field</span>. During consolidated undrained multi-stage triaxial tests on core plugs cut normal to bedding, V pv increases monotonically with increasing effective stress and V s1 behaves similarly although with some scatter. V ph and V sh remain constant initially but then decrease within each stage of the multi-stage test, although <span class="hlt">velocity</span> from stage to stage at any given differential stress increases. This has the impact of decreasing both P-wave (ɛ) and S-wave anisotropy (γ) through application of differential stress within each loading stage. However, increasing the magnitude of an isotropic stress <span class="hlt">field</span> has little effect on the <span class="hlt">velocity</span> anisotropies. The intrinsic anisotropy of the shale remains reasonably high at the highest confining pressures. The results indicate the magnitude and orientation of the stress anisotropy with respect to the shale microfabric has a significant impact on the <span class="hlt">velocity</span> response to changing stress <span class="hlt">fields</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120004056','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120004056"><span><span class="hlt">Velocity</span> Vector <span class="hlt">Field</span> Visualization of Flow in Liquid Acquisition Device Channel</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>McQuillen, John B.; Chao, David F.; Hall, Nancy R.; Zhang, Nengli</p> <p>2012-01-01</p> <p>A capillary flow liquid acquisition device (LAD) for cryogenic propellants has been developed and tested in NASA Glenn Research Center to meet the requirements of transferring cryogenic liquid propellants from storage tanks to an engine in reduced gravity environments. The prototypical mesh screen channel LAD was fabricated with a mesh screen, covering a rectangular flow channel with a cylindrical outlet tube, and was tested with liquid oxygen (LOX). In order to better understand the performance in various gravity environments and orientations at different liquid submersion depths of the screen channel LAD, a series of computational fluid dynamics (CFD) simulations of LOX flow through the LAD screen channel was undertaken. The resulting <span class="hlt">velocity</span> vector <span class="hlt">field</span> visualization for the flow in the channel has been used to reveal the gravity effects on the flow in the screen channel.</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/2017AGUFM.S54A..02C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.S54A..02C"><span>Simultaneous inversion of seismic <span class="hlt">velocity</span> and moment tensor using elastic-waveform inversion of microseismic data: Application to the Aneth CO2-EOR <span class="hlt">field</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, Y.; Huang, L.</p> <p>2017-12-01</p> <p>Moment tensors are key parameters for characterizing CO2-injection-induced microseismic events. Elastic-waveform inversion has the potential to providing accurate results of moment tensors. Microseismic waveforms contains information of source moment tensors and the wave propagation <span class="hlt">velocity</span> along the wavepaths. We develop an elastic-waveform inversion method to jointly invert the seismic <span class="hlt">velocity</span> model and moment tensor. We first use our adaptive moment-tensor joint inversion method to estimate moment tensors of microseismic events. Our adaptive moment-tensor inversion method jointly inverts multiple microseismic events with similar waveforms within a cluster to reduce inversion uncertainty for microseismic data recorded using a single borehole geophone array. We use this inversion result as the initial model for our elastic-waveform inversion to minimize the cross-correlated-based data misfit between <span class="hlt">observed</span> data and synthetic data. We verify our method using synthetic microseismic data and obtain improved results of both moment tensors and seismic <span class="hlt">velocity</span> model. We apply our new inversion method to microseismic data acquired at a CO2-enhanced oil recovery <span class="hlt">field</span> in Aneth, Utah, using a single borehole geophone array. The results demonstrate that our new inversion method significantly reduces the data misfit compared to the conventional ray-theory-based moment-tensor inversion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.G13A0994O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.G13A0994O"><span>Slip Rates of Main Active Fault Zones Through Turkey Inferred From 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>Ozener, H.; Aktug, B.; Dogru, A.; Tasci, L.; Acar, M.; Emre, O.; Yilmaz, O.; Turgut, B.; Halicioglu, K.; Sabuncu, A.; Bal, O.; Eraslan, A.</p> <p>2015-12-01</p> <p>Active Fault Map of Turkey was revised and published by General Directorate of Mineral Research and Exploration in 2012. This map reveals that there are about 500 faults can generate earthquakes.In order to understand the earthquake potential of these faults, it is needed to determine the slip rates. Although many regional and local studies were performed in the past, the slip rates of the active faults in Turkey have not been determined. In this study, the block modelling, which is the most common method to produce slip rates, will be done. GPS <span class="hlt">velocities</span> required for block modeling is being compiled from the published studies and the raw data provided then <span class="hlt">velocity</span> <span class="hlt">field</span> is combined. To form a homogeneous <span class="hlt">velocity</span> <span class="hlt">field</span>, different stochastic models will be used and the optimal <span class="hlt">velocity</span> <span class="hlt">field</span> will be achieved. In literature, GPS site <span class="hlt">velocities</span>, which are computed for different purposes and published, are combined globally and this combined <span class="hlt">velocity</span> <span class="hlt">field</span> are used in the analysis of strain accumulation. It is also aimed to develop optimal stochastic models to combine the <span class="hlt">velocity</span> data. Real time, survey mode and published GPS <span class="hlt">observations</span> is being combined in this study. We also perform new GPS <span class="hlt">observations</span>. Furthermore, micro blocks and main fault zones from Active Fault Map Turkey will be determined and homogeneous <span class="hlt">velocity</span> <span class="hlt">field</span> will be used to infer slip rates of these active faults. Here, we present the result of first year of the study. This study is being supported by THE SCIENTIFIC AND TECHNOLOGICAL RESEARCH COUNCIL OF TURKEY (TUBITAK)-CAYDAG with grant no. 113Y430.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/212191-dean-vortices-wall-flux-curved-channel-membrane-system-velocity-field','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/212191-dean-vortices-wall-flux-curved-channel-membrane-system-velocity-field"><span>Dean vortices with wall flux in a curved channel membrane system. 2: The <span class="hlt">velocity</span> <span class="hlt">field</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>Chung, K.Y.; Brewster, M.E.; Belfort, G.</p> <p>1996-02-01</p> <p>The <span class="hlt">velocity</span> and pressure <span class="hlt">fields</span> and the effect of wall flux on these <span class="hlt">fields</span> in a spiral channel are presented. As fluid flows inward through a spiral channel with constant gap and permeable walls, the streamwise flux decreases while the curvature increases. Thus, by balancing the stabilizing effect of wall suction with the destabilizing effect of increasing curvature, established vortices can be maintained along the spiral channel. This approach is used to prescribe spiral geometries with different wall fluxes. Using a weakly nonlinear stability analysis, the influence of wall flux on the characteristics of Dean vortices is obtained. The criticalmore » Dean number is reduced when suction is through the inner wall only, is slightly reduced when suction is equal through both walls, and is increased when suction is through the outer wall only. The magnitude of change is proportional to a ratio of small numbers that measures the importance of the effect of curvature. In membrane filtration applications the wall flux is typically 2 to 5 orders of magnitude less than the streamwise flow. If the radius of curvature of the channel is of the order of 100 times the channel gap, the effect on the critical Dean number is within 2% of the no-wall flux case. If the radius of curvature is sufficiently large, however, it is possible to <span class="hlt">observe</span> effects on the critical Dean number that approach O(1) in magnitude for certain parameter ranges.« 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_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/2015APS..DFDE27003H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..DFDE27003H"><span>Development of a 3-wire probe for the simultaneous measurement of turbulent <span class="hlt">velocity</span>, concentration and temperature <span class="hlt">fields</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hewes, Alaïs; Mydlarski, Laurent</p> <p>2015-11-01</p> <p>The present work focuses on the design and optimization of a probe used to simultaneously measure the <span class="hlt">velocity</span>, concentration and temperature <span class="hlt">fields</span> in a turbulent jet. The underlying principles of this sensor are based in thermal-anemometry techniques, and the design of this 3-wire probe builds off the previous work of Sirivat and Warhaft, J. Fluid Mech., 1982. In the first part of this study, the effect of different overheat ratios in the first two wires (called the ``interference'' or ``Way-Libby'' probe - used to infer <span class="hlt">velocity</span> and concentration) are investigated. Of particular interest is their effect on the quality of the resulting calibration, as well as the measured <span class="hlt">velocity</span> and concentration data. Four different overheat ratio pairs for the two wires comprising the interference probe are studied. In the second part of this work, a third wire, capable of detecting temperature fluctuations, is added to the 3-wire probe. The optimal configuration of this probe, including wire type and overheat ratio for the third wire, is studied and the simultaneously-measured <span class="hlt">velocity</span>, concentration, and temperature data (e.g. spectra, PDFs) for different probe configurations are presented. Supported by the Natural Sciences and Engineering Research Council of Canada (Grant 217184).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040082225&hterms=Gradient&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DGradient','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040082225&hterms=Gradient&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DGradient"><span>Pickup Ion <span class="hlt">Velocity</span> Distributions at Titan: Effects of Spatial Gradients</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hartle, R. E.; Sittler, E. C.</p> <p>2004-01-01</p> <p>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 <span class="hlt">field</span> and background plasma. When the pickup ion gyroradius is less than the source gas scale height, the pickup ion <span class="hlt">velocity</span> 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 <span class="hlt">field</span> and the flow <span class="hlt">velocity</span>. 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 <span class="hlt">velocity</span> 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 <span class="hlt">observed</span> from Voyager 1 , but should be <span class="hlt">observable</span> by more sensitive instruments on future spacecraft, including Cassini. The finite gyroradius effects in the pickup ion <span class="hlt">velocity</span> distributions are studied by including in the analysis the possible range of spatial structures in the neutral exosphere and background plasma.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007JKAS...40..165S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007JKAS...40..165S"><span>Remote Numerical Simulations of the Interaction of High <span class="hlt">Velocity</span> Clouds with Random Magnetic <span class="hlt">Fields</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Santillan, Alfredo; Hernandez--Cervantes, Liliana; Gonzalez--Ponce, Alejandro; Kim, Jongsoo</p> <p></p> <p>The numerical simulations associated with the interaction of High <span class="hlt">Velocity</span> Clouds (HVC) with the Magnetized Galactic Interstellar Medium (ISM) are a powerful tool to describe the evolution of the interaction of these objects in our Galaxy. In this work we present a new project referred to as Theoretical Virtual i Observatories. It is oriented toward to perform numerical simulations in real time through a Web page. This is a powerful astrophysical computational tool that consists of an intuitive graphical user interface (GUI) and a database produced by numerical calculations. In this Website the user can make use of the existing numerical simulations from the database or run a new simulation introducing initial conditions such as temperatures, densities, <span class="hlt">velocities</span>, and magnetic <span class="hlt">field</span> intensities for both the ISM and HVC. The prototype is programmed using Linux, Apache, MySQL, and PHP (LAMP), based on the open source philosophy. All simulations were performed with the MHD code ZEUS-3D, which solves the ideal MHD equations by finite differences on a fixed Eulerian mesh. Finally, we present typical results that can be obtained with this tool.</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.474..522O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.474..522O"><span>The connection between the peaks in <span class="hlt">velocity</span> dispersion and star-forming clumps of turbulent galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Oliva-Altamirano, P.; Fisher, D. B.; Glazebrook, K.; Wisnioski, E.; Bekiaris, G.; Bassett, R.; Obreschkow, D.; Abraham, R.</p> <p>2018-02-01</p> <p>We present Keck/OSIRIS adaptive optics <span class="hlt">observations</span> with 150-400 pc spatial sampling of 7 turbulent, clumpy disc galaxies from the DYNAMO sample ($0.07<z<0.2$). DYNAMO galaxies have previously been shown to be well matched in properties to main sequence galaxies at $z\\sim1.5$. Integral <span class="hlt">field</span> spectroscopy <span class="hlt">observations</span> using adaptive optics are subject to a number of systematics including a variable PSF and spatial sampling, which we account for in our analysis. We present gas <span class="hlt">velocity</span> dispersion maps corrected for these effects, and confirm that DYNAMO galaxies do have high gas <span class="hlt">velocity</span> dispersion ($\\sigma=40-80$\\kms), even at high spatial sampling. We find statistically significant structure in 6 out of 7 galaxies. The most common distance between the peaks in <span class="hlt">velocity</span> dispersion and emission line peaks is $\\sim0.5$~kpc, we note this is very similar to the average size of a clump measured with HST H$\\alpha$ maps. This could suggest that the peaks in <span class="hlt">velocity</span> dispersion in clumpy galaxies likely arise due to some interaction between the clump and the surrounding ISM of the galaxy, though our <span class="hlt">observations</span> cannot distinguish between outflows, inflows or <span class="hlt">velocity</span> shear. <span class="hlt">Observations</span> covering a wider area of the galaxies will be needed to confirm this result.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ACP....18.6761D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ACP....18.6761D"><span>Quantifying the effect of aerosol on vertical <span class="hlt">velocity</span> and effective terminal <span class="hlt">velocity</span> in warm convective clouds</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dagan, Guy; Koren, Ilan; Altaratz, Orit</p> <p>2018-05-01</p> <p>Better representation of cloud-aerosol interactions is crucial for an improved understanding of natural and anthropogenic effects on climate. Recent studies have shown that the overall aerosol effect on warm convective clouds is non-monotonic. Here, we reduce the system's dimensions to its center of gravity (COG), enabling distillation and simplification of the overall trend and its temporal evolution. Within the COG framework, we show that the aerosol effects are nicely reflected by the interplay of the system's characteristic vertical <span class="hlt">velocities</span>, namely the updraft (w) and the effective terminal <span class="hlt">velocity</span> (η). The system's vertical <span class="hlt">velocities</span> can be regarded as a sensitive measure for the evolution of the overall trends with time. Using a bin-microphysics cloud-scale model, we analyze and follow the trends of the aerosol effect on the magnitude and timing of w and η, and therefore the overall vertical COG <span class="hlt">velocity</span>. Large eddy simulation (LES) model runs are used to upscale the analyzed trends to the cloud-<span class="hlt">field</span> scale and study how the aerosol effects on the temporal evolution of the <span class="hlt">field</span>'s thermodynamic properties are reflected by the interplay between the two <span class="hlt">velocities</span>. Our results suggest that aerosol effects on air vertical motion and droplet mobility imply an effect on the way in which water is distributed along the atmospheric column. Moreover, the interplay between w and η predicts the overall trend of the <span class="hlt">field</span>'s thermodynamic instability. These factors have an important effect on the local energy balance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.H43H1051L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.H43H1051L"><span>The Continuous Monitoring of Flash Flood <span class="hlt">Velocity</span> <span class="hlt">Field</span> based on an Automated LSPIV System</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.; Ran, Q.; Liao, Q.</p> <p>2014-12-01</p> <p>Large-scale particle image velocimetry (LSPIV) is a non-intrusive tool for flow <span class="hlt">velocity</span> <span class="hlt">field</span> measurement and has more advantages against traditional techniques, with its applications on river, lake and ocean, especially under extreme conditions. An automated LSPIV system is presented in this study, which can be easily set up and executed for continuous monitoring of flash flood. The experiment site is Longchi village, Sichuan Province, where 8.0 magnitude earthquake occurred in 2008 and debris flow happens every year since then. The interest of area is about 30m*40m of the channel which has been heavily destroyed by debris flow. Series of videos obtained during the flood season indicates that flood outbreaks after rainstorm just for several hours. Measurement is complete without being influenced by this extreme weather condition and results are more reliable and accurate due to high soil concentration. Compared with direct measurement by impellor flow meter, we validated that LSPIV works well at mountain stream, with index of 6.7% (Average Relative Error) and 95% (Nash-Sutcliffe Coefficient). On Jun 26, the maximum flood surface <span class="hlt">velocity</span> reached 4.26 m/s, and the discharge based on <span class="hlt">velocity</span>-area method was also decided. Overall, this system is safe, non-contact and can be adjusted according to our requirement flexibly. We can get valuable data of flood which is scarce before, which will make a great contribution to the analysis of flood and debris flow mechanism.</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> <span class="hlt">field</span> does not contain discontinuities or steep gradients 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/2016AGUFMSH43B2575P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSH43B2575P"><span>Analysis of <span class="hlt">velocity</span> and magnetic <span class="hlt">field</span> fluctuations from simulated Solar Probe Plus measurements: Interpretation and predictions.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Perez, J. C.; Chandran, B. D. G.</p> <p>2016-12-01</p> <p>As Solar Probe Plus (SPP) explores the near-Sun environment, our ability to obtain meaningful interpretation of in-situ measurements faces two significant challenges. The first challenge is that the Taylor Hypothesis (TH), which is normally used in the interpretation of existing spacecraft data, breaks down at the low heliocentric distances that SPP mission will explore. The second challenge is our limited understanding of turbulence in this region, largely due to the theoretical and numerical difficulties in modeling this problem. In this work we present recent progress towards overcoming these challenges using high-resolution numerical simulations of Alfvenic turbulence in the inner heliosphere. We fly virtual SPP spacecraft in the simulation domain to obtain single-point measurements of the <span class="hlt">velocity</span> and magnetic <span class="hlt">field</span> fluctuations at several radial locations relevant to SPP. We use these virtual measurements to 1) validate a recently introduced modified TH that allows one to recover the spatial structure of the dominant (outward-propagating) Alfvenic fluctuations, of the kind SPP will encounter; and 2) to compare these virtual <span class="hlt">observations</span> with our most recent phenomenological models of reflection-driven Alfven turbulence.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JGRA..119.2400P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JGRA..119.2400P"><span>Generation of temperature anisotropy for alpha particle <span class="hlt">velocity</span> distributions in solar wind at 0.3 AU: Vlasov simulations and Helios <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>Perrone, D.; Bourouaine, S.; Valentini, F.; Marsch, E.; Veltri, P.</p> <p>2014-04-01</p> <p>Solar wind "in situ" measurements from the Helios spacecraft in regions of the Heliosphere close to the Sun (˜0.3 AU), at which typical values of the proton plasma beta are <span class="hlt">observed</span> to be lower than unity, show that the alpha particle distribution functions depart from the equilibrium Maxwellian configuration, displaying significant elongations in the direction perpendicular to the background magnetic <span class="hlt">field</span>. In the present work, we made use of multi-ion hybrid Vlasov-Maxwell simulations to provide theoretical support and interpretation to the empirical evidences above. Our numerical results show that, at variance with the case of βp≃1 discussed in Perrone et al. (2011), for βp=0.1 the turbulent cascade in the direction parallel to the ambient magnetic <span class="hlt">field</span> is not efficient in transferring energy toward scales shorter than the proton inertial length. Moreover, our numerical analysis provides new insights for the theoretical interpretation of the empirical evidences obtained from the Helios spacecraft, concerning the generation of temperature anisotropy in the particle <span class="hlt">velocity</span> distributions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013HEAD...1312205W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013HEAD...1312205W"><span>Magnetized Collisionless Shock Studies Using High <span class="hlt">Velocity</span> Plasmoids</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Weber, Thomas; Intrator, T.</p> <p>2013-04-01</p> <p>Magnetized collisionless shocks are ubiquitous throughout the cosmos and are <span class="hlt">observed</span> to accelerate particles to relativistic <span class="hlt">velocities</span>, amplify magnetic <span class="hlt">fields</span>, transport energy, and create non-thermal distributions. They exhibit transitional scale lengths much shorter than the collisional mean free path and are mediated by collective interactions rather than Coulomb collisions. The Magnetized Shock Experiment (MSX) leverages advances in <span class="hlt">Field</span> Reversed Configuration (FRC) plasmoid formation and acceleration to produce highly supersonic and super-Alfvénic supercritical shocks with pre-existing magnetic <span class="hlt">field</span> at perpendicular, parallel or oblique angles to the direction of propagation. Adjustable shock speed, density, and magnetic <span class="hlt">field</span> provide unique access to a range of parameter space relevant to a variety of naturally occurring shocks. This effort examines experimentally, analytically, and numerically the physics of collisionless shock formation, structure, and kinetic effects in a laboratory setting and draw comparisons between experimental data and astronomical <span class="hlt">observations</span>. Supported by DOE Office of Fusion Energy Sciences and National Nuclear Security Administration under LANS contract DE-AC52-06NA25369 Approved for Public Release: LA-UR-12-22886</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.ncbi.nlm.nih.gov/pubmed/16383790','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16383790"><span>Statistical <span class="hlt">field</span> estimators for multiscale simulations.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Eapen, Jacob; Li, Ju; Yip, Sidney</p> <p>2005-11-01</p> <p>We present a systematic approach for generating smooth and accurate <span class="hlt">fields</span> from particle simulation data using the notions of statistical inference. As an extension to a parametric representation based on the maximum likelihood technique previously developed for <span class="hlt">velocity</span> and temperature <span class="hlt">fields</span>, a nonparametric estimator based on the principle of maximum entropy is proposed for particle density and stress <span class="hlt">fields</span>. Both estimators are applied to represent molecular dynamics data on shear-driven flow in an enclosure which exhibits a high degree of nonlinear characteristics. We show that the present density estimator is a significant improvement over ad hoc bin averaging and is also free of systematic boundary artifacts that appear in the method of smoothing kernel estimates. Similarly, the <span class="hlt">velocity</span> <span class="hlt">fields</span> generated by the maximum likelihood estimator do not show any edge effects that can be erroneously interpreted as slip at the wall. For low Reynolds numbers, the <span class="hlt">velocity</span> <span class="hlt">fields</span> and streamlines generated by the present estimator are benchmarked against Newtonian continuum calculations. For shear <span class="hlt">velocities</span> that are a significant fraction of the thermal speed, we <span class="hlt">observe</span> a form of shear localization that is induced by the confining boundary.</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/2016AGUOS.A33A..04B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUOS.A33A..04B"><span><span class="hlt">Observations</span> of whitecaps during HiWinGS, their dependence on wave <span class="hlt">field</span>, and relation to gas transfer <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>Brumer, S. E.; Zappa, C. J.; Fairall, C. W.; Blomquist, B.; Brooks, I. M.; Tamura, H.; Yang, M.; Huebert, B. J.</p> <p>2016-02-01</p> <p>The High Wind Gas exchange Study (HiWinGS) presents the unique opportunity to gain new insights on the poorly understood aspects of air-sea interaction under high winds. The HiWinGS cruise took place in the North Atlantic during October and November 2013. Wind speeds exceeded 15 m s-1 25% of the time, including 48 hrs with U10 > 20 m s-1. Continuous measurements of turbulent fluxes of heat, momentum, and gas were taken from the bow of the R/V Knorr. Visible imagery was acquired from the port and starboard side of the flying bridge during daylight hours at 20Hz and directional wave spectra were obtained when on station from a wave rider buoy. Additional wave <span class="hlt">field</span> statistics were computed from a laser altimeter as well as from a Wavewatch III hindcast. Taking advantage of the range of physical forcing and wave conditions sampled during HiWinGS, we investigate how the fractional whitecap coverage (W) and gas transfer <span class="hlt">velocity</span> (K) vary with sea state. We distinguish between windseas and swell based on a separation algorithm applied to directional wave spectra, allowing contrasting pure windseas to swell dominated periods. For mixed seas, system alignment is considered when interpreting results. The four gases sampled during HiWinGS ranged from being mostly waterside controlled to almost entirely airside controlled. While bubble-mediated transfer appears to be small for moderately soluble gases like DMS, the importance of wave breaking turbulence transport has yet to be determined for all gases regardless of their solubility. This will be addressed by correlating measured K to estimates of active whitecap fraction (WA) and turbulent kinetic energy dissipation rate (ɛ). WA and ɛ are estimated from moments of the breaking crest length distribution derived from the imagery, focusing on young seas, when it is likely that large-scale breaking waves (i.e., whitecapping) will dominate the ɛ.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1993PhDT........25D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1993PhDT........25D"><span>A finite element solution to conjugated heat transfer in tissue using magnetic resonance angiography to measure the in vitro <span class="hlt">velocity</span> <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dutton, Andrew William</p> <p>1993-12-01</p> <p>A combined numerical and experimental system for tissue heat transfer analysis was developed. The goal was to develop an integrated set of tools for studying the problem of providing accurate temperature estimation for use in hyperthermia treatment planning in a clinical environment. The completed system combines (1) Magnetic Resonance Angiography (MRA) to non-destructively measure the <span class="hlt">velocity</span> <span class="hlt">field</span> in situ, (2) the Streamwise Upwind Petrov-Galerkin finite element solution to the 3D steady state convective energy equation (CEE), (3) a medical image based automatic 3D mesh generator, and (4) a Gaussian type estimator to determine unknown thermal model parameters such as thermal conductivity, blood perfusion, and blood <span class="hlt">velocities</span> from measured temperature data. The system was capable of using any combination of three thermal models (1) the Convective Energy Equation (CEE), (2) the Bioheat Transfer Equation (BHTE), and (3) the Effective Thermal Conductivity Equation (ETCE) Incorporation of the theoretically correct CEE was a significant theoretical advance over approximate models made possible by the use of MRA to directly measure the 3D <span class="hlt">velocity</span> <span class="hlt">field</span> in situ. Experiments were carried out in a perfused alcohol fixed canine liver with hyperthermia induced through scanned focused ultrasound <span class="hlt">Velocity</span> <span class="hlt">fields</span> were measured using Phase Contrast Angiography. The complete system was then used to (1) develop a 3D finite element model based upon user traced outlines over a series of MR images of the liver and (2) simulate temperatures at steady state using the CEE, BHTE, and ETCE thermal models in conjunction with the gauss estimator. Results of using the system on an in vitro liver preparation indicate the need for improved accuracy in the MRA scans and accurate spatial registration between the thermocouple junctions, the measured <span class="hlt">velocity</span> <span class="hlt">field</span>, and the scanned ultrasound power No individual thermal model was able to meet the desired accuracy of 0.5 deg C, the resolution</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012APS..DFDL14008V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012APS..DFDL14008V"><span>On the extraction of pressure <span class="hlt">fields</span> from PIV <span class="hlt">velocity</span> measurements in turbines</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Villegas, Arturo; Diez, Fancisco J.</p> <p>2012-11-01</p> <p>In this study, the pressure <span class="hlt">field</span> for a water turbine is derived from particle image velocimetry (PIV) measurements. Measurements are performed in a recirculating water channel facility. The PIV measurements include calculating the tangential and axial forces applied to the turbine by solving the integral momentum equation around the airfoil. The results are compared with the forces obtained from the Blade Element Momentum theory (BEMT). Forces are calculated by using three different methods. In the first method, the pressure <span class="hlt">fields</span> are obtained from PIV <span class="hlt">velocity</span> <span class="hlt">fields</span> by solving the Poisson equation. The boundary conditions are obtained from the Navier-Stokes momentum equations. In the second method, the pressure at the boundaries is determined by spatial integration of the pressure gradients along the boundaries. In the third method, applicable only to incompressible, inviscid, irrotational, and steady flow, the pressure is calculated using the Bernoulli equation. This approximated pressure is known to be accurate far from the airfoil and outside of the wake for steady flows. Additionally, the pressure is used to solve for the force from the integral momentum equation on the blade. From the three methods proposed to solve for pressure and forces from PIV measurements, the first one, which is solved by using the Poisson equation, provides the best match to the BEM theory calculations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15898628','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15898628"><span>Simultaneous measurement of acoustic and streaming <span class="hlt">velocities</span> in a standing wave using laser Doppler anemometry.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Thompson, Michael W; Atchley, Anthony A</p> <p>2005-04-01</p> <p>Laser Doppler anemometry (LDA) with burst spectrum analysis (BSA) is used to study the acoustic streaming generated in a cylindrical standing-wave resonator filled with air. The air column is driven sinusoidally at a frequency of approximately 310 Hz and the resultant acoustic-<span class="hlt">velocity</span> amplitudes are less than 1.3 m/s at the <span class="hlt">velocity</span> antinodes. The axial component of fluid <span class="hlt">velocity</span> is measured along the resonator axis, across the diameter, and as a function of acoustic amplitude. The <span class="hlt">velocity</span> signals are postprocessed using the Fourier averaging method [Sonnenberger et al., Exp. Fluids 28, 217-224 (2000)]. Equations are derived for determining the uncertainties in the resultant Fourier coefficients. The time-averaged <span class="hlt">velocity</span>-signal components are seen to be contaminated by significant errors due to the LDA/BSA system. In order to avoid these errors, the Lagrangian streaming <span class="hlt">velocities</span> are determined using the time-harmonic signal components and the arrival times of the <span class="hlt">velocity</span> samples. The <span class="hlt">observed</span> Lagrangian streaming <span class="hlt">velocities</span> are consistent with Rott's theory [N. Rott, Z. Angew. Math. Phys. 25, 417-421 (1974)], indicating that the dependence of viscosity on temperature is important. The onset of streaming is <span class="hlt">observed</span> to occur within approximately 5 s after switching on the acoustic <span class="hlt">field</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhDT.......265R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhDT.......265R"><span>A seismic reflection <span class="hlt">velocity</span> study of a Mississippian mud-mound in the Illinois basin</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ranaweera, Chamila Kumari</p> <p></p> <p>. The 1-D synthetics were used by Cory Cantrell of Royal Drilling and Producing Company to identify various reflections on the seismic records. Seismic data was compared with the modeled synthetic seismograms to identify what appears to be a carbonate mud-mound within the Aden study area. No mud-mounds have been previously found in the Aden oil <span class="hlt">field</span>. Average and interval <span class="hlt">velocities</span> obtained from the geophysical logs from the wells drilled in the Aden area was compared with the same type of well <span class="hlt">velocities</span> from the Broughton known mud-mound area to <span class="hlt">observe</span> the significance of <span class="hlt">velocity</span> variation related to the un-known mud-mound in the Aden study area. The results of the <span class="hlt">velocity</span> study shows a similar trends in the wells from both areas and are higher at the bottom of the wells. Another approach was used to <span class="hlt">observe</span> the variation of root mean square <span class="hlt">velocities</span> calculated from the sonic log from the well <span class="hlt">velocity</span> from the Aden area and the stacking <span class="hlt">velocities</span> obtained from the seismic data adjacent to the well.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29021095','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29021095"><span>The role of wind <span class="hlt">field</span> induced flow <span class="hlt">velocities</span> in destratification and hypoxia reduction at Meiling Bay of large shallow Lake Taihu, China.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jalil, Abdul; Li, Yiping; Du, Wei; Wang, Wencai; Wang, Jianwei; Gao, Xiaomeng; Khan, Hafiz Osama Sarwar; Pan, Baozhu; Acharya, Kumud</p> <p>2018-01-01</p> <p>Wind induced flow <span class="hlt">velocity</span> patterns and associated thermal destratification can drive to hypoxia reduction in large shallow lakes. The effects of wind induced hydrodynamic changes on destratification and hypoxia reduction were investigated at the Meiling bay (N 31° 22' 56.4″, E 120° 9' 38.3″) of Lake Taihu, China. Vertical flow <span class="hlt">velocity</span> profile analysis showed surface flow <span class="hlt">velocities</span> consistency with the wind <span class="hlt">field</span> and lower flow <span class="hlt">velocity</span> profiles were also consistent (but with delay response time) when the wind speed was higher than 6.2 m/s. Wind <span class="hlt">field</span> and temperature found the control parameters for hypoxia reduction and for water quality conditions at the surface and bottom profiles of lake. The critical temperature for hypoxia reduction at the surface and the bottom profile was ≤24.1C° (below which hypoxic conditions were found reduced). Strong prevailing wind <span class="hlt">field</span> (onshore wind directions ESE, SE, SSE and E, wind speed ranges of 2.4-9.1 m/s) reduced the temperature (22C° to 24.1C°) caused reduction of hypoxia at the near surface with a rise in water levels whereas, low to medium prevailing wind <span class="hlt">field</span> did not supported destratification which increased temperature resulting in increased hypoxia. Non-prevailing wind directions (offshore) were not found supportive for the reduction of hypoxia in study area due to less variable wind <span class="hlt">field</span>. Daytime wind <span class="hlt">field</span> found more variable (as compared to night time) which increased the thermal destratification during daytime and found supportive for destratification and hypoxia reduction. The second order exponential correlation found between surface temperature and Chlorophyll-a (R 2 : 0.2858, Adjusted R-square: 0.2144 RMSE: 4.395), Dissolved Oxygen (R 2 : 0.596, Adjusted R-square: 0.5942, RMSE: 0.3042) concentrations. The findings of the present study reveal the driving mechanism of wind induced thermal destratification and hypoxic conditions, which may further help to evaluate the wind role in eutrophication</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/2009AGUFM.S21B1721Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFM.S21B1721Z"><span>Strain-dependent Damage Evolution and <span class="hlt">Velocity</span> Reduction in Fault Zones Induced by Earthquake Rupture</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhong, J.; Duan, B.</p> <p>2009-12-01</p> <p>Low-<span class="hlt">velocity</span> fault zones (LVFZs) with reduced seismic <span class="hlt">velocities</span> relative to the surrounding wall rocks are widely <span class="hlt">observed</span> around active faults. The presence of such a zone will affect rupture propagation, near-<span class="hlt">field</span> ground motion, and off-fault damage in subsequent earth-quakes. In this study, we quantify the reduction of seismic <span class="hlt">velocities</span> caused by dynamic rup-ture on a 2D planar fault surrounded by a low-<span class="hlt">velocity</span> fault zone. First, we implement the damage rheology (Lyakhovsky et al. 1997) in EQdyna (Duan and Oglesby 2006), an explicit dynamic finite element code. We further extend this damage rheology model to include the dependence of strains on crack density. Then, we quantify off-fault continuum damage distribution and <span class="hlt">velocity</span> reduction induced by earthquake rupture with the presence of a preexisting LVFZ. We find that the presence of a LVFZ affects the tempo-spatial distribu-tions of off-fault damage. Because lack of constraint in some damage parameters, we further investigate the relationship between <span class="hlt">velocity</span> reduction and these damage prameters by a large suite of numerical simulations. Slip <span class="hlt">velocity</span>, slip, and near-<span class="hlt">field</span> ground motions computed from damage rheology are also compared with those from off-fault elastic or elastoplastic responses. We find that the reduction in elastic moduli during dynamic rupture has profound impact on these quantities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ThEng..64..802D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ThEng..64..802D"><span>Experimental study of the possibility of reducing the resistance and unevenness of output <span class="hlt">field</span> of <span class="hlt">velocities</span> in flat diffuser channels with large opening angles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dmitriev, S. S.; Vasil'ev, K. E.; Mokhamed, S. M. S. O.; Gusev, A. A.; Barbashin, A. V.</p> <p>2017-11-01</p> <p>In modern combined cycle gas turbines (CCGT), when designing the reducers from the output diffuser of a gas turbine to a boiler-utilizer, wide-angle diffusers are used, in which practically from the input a flow separation and transition to jet stream regime occurs. In such channels, the energy loss in the <span class="hlt">field</span> of <span class="hlt">velocities</span> sharply rise and the <span class="hlt">field</span> of <span class="hlt">velocities</span> in the output from them is characterized by considerable unevenness that worsens the heat transfer process in the first by motion tube bundles of the boiler-utilizer. The results of experimental research of the method for reducing the energy loss and alignment of the <span class="hlt">field</span> of <span class="hlt">velocities</span> at the output from a flat asymmetrical diffuser channel with one deflecting wall with the opening angle of 40° by means of placing inside the channel the flat plate parallel to the deflecting wall are presented in the paper. It is revealed that, at this placement of the plate in the channel, it has a chance to reduce the energy loss by 20%, considerably align the output <span class="hlt">field</span> of <span class="hlt">velocities</span>, and decrease the dynamic loads on the walls in the output cross-section. The studied method of resistance reduction and alignment of the <span class="hlt">fields</span> of <span class="hlt">velocities</span> in the flat diffuser channels was used for optimization of the reducer from the output diffuser of the gas turbine to the boiler-utilizer of CCGT of PGU-450T type of Kaliningrad Thermal Power Plant-2. The obtained results are evidence that the configuration of the reducer installed in the PGU-450T of Kaliningrad Thermal Power Plant-2 is not optimal. It follows also from the obtained data that working-off the reducer should be necessarily conducted by the test results of the channel consisting of the model of reducer with the model of boiler-utilizer installed behind it. Application of the method of alignment of output <span class="hlt">field</span> of <span class="hlt">velocities</span> and reducing the resistance in the wide-angle diffusers investigated in the work made it possible—when using the known model of diffusion</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://pubs.er.usgs.gov/publication/70180072','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70180072"><span>Relationships among seismic <span class="hlt">velocity</span>, metamorphism, and seismic and aseismic fault slip in the Salton Sea Geothermal <span class="hlt">Field</span> region</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>McGuire, Jeffrey J.; Lohman, Rowena B.; Catchings, Rufus D.; Rymer, Michael J.; Goldman, Mark R.</p> <p>2015-01-01</p> <p>The Salton Sea Geothermal <span class="hlt">Field</span> is one of the most geothermally and seismically active areas in California and presents an opportunity to study the effect of high-temperature metamorphism on the properties of seismogenic faults. The area includes numerous active tectonic faults that have recently been imaged with active source seismic reflection and refraction. We utilize the active source surveys, along with the abundant microseismicity data from a dense borehole seismic network, to image the 3-D variations in seismic <span class="hlt">velocity</span> in the upper 5 km of the crust. There are strong <span class="hlt">velocity</span> variations, up to ~30%, that correlate spatially with the distribution of shallow heat flow patterns. The combination of hydrothermal circulation and high-temperature contact metamorphism has significantly altered the shallow sandstone sedimentary layers within the geothermal <span class="hlt">field</span> to denser, more feldspathic, rock with higher P wave <span class="hlt">velocity</span>, as is seen in the numerous exploration wells within the <span class="hlt">field</span>. This alteration appears to have a first-order effect on the frictional stability of shallow faults. In 2005, a large earthquake swarm and deformation event occurred. Analysis of interferometric synthetic aperture radar data and earthquake relocations indicates that the shallow aseismic fault creep that occurred in 2005 was localized on the Kalin fault system that lies just outside the region of high-temperature metamorphism. In contrast, the earthquake swarm, which includes all of the M > 4 earthquakes to have occurred within the Salton Sea Geothermal <span class="hlt">Field</span> in the last 15 years, ruptured the Main Central Fault (MCF) system that is localized in the heart of the geothermal anomaly. The background microseismicity induced by the geothermal operations is also concentrated in the high-temperature regions in the vicinity of operational wells. However, while this microseismicity occurs over a few kilometer scale region, much of it is clustered in earthquake swarms that last from</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.V33E3177W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.V33E3177W"><span>Preliminary Seismic <span class="hlt">Velocity</span> Structure Results from Ambient Noise and Teleseismic Tomography: Laguna del Maule Volcanic <span class="hlt">Field</span>, Chile</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wespestad, C.; Thurber, C. H.; Zeng, X.; Bennington, N. L.; Cardona, C.; Singer, B. S.</p> <p>2016-12-01</p> <p>Laguna del Maule Volcanic <span class="hlt">Field</span> is a large, restless, rhyolitic system in the Southern Andes that is being heavily studied through several methods, including seismology, by a collaborative team of research institutions. A temporary array of 52 seismometers from OVDAS (the Southern Andean Volcano Observatory), PASSCAL (Portable Array Seismic Studies of the Continental Lithosphere), and the University of Wisconsin-Madison was used to collect the 1.3 years worth of data for this preliminary study. Ambient noise tomography uses surface wave dispersion data obtained from noise correlation functions (NCFs) between pairs of seismic stations, with one of each pair acting as a virtual source, in order to image the <span class="hlt">velocity</span> structure in 3-D. NCFs were computed for hour-long time windows, and the final NCFs were obtained with phase-weighted stacking. The Frequency-Time Analysis technique was then utilized to measure group <span class="hlt">velocity</span> between station pairs. NCFs were also analyzed to detect temporal changes in seismic <span class="hlt">velocity</span> related to magmatic activity at the volcano. With the surface wave data from ambient noise, our small array aperture limits our modeling to the upper crust, so we employed teleseismic tomography to study deeper structures. For picking teleseismic arrivals, we tested two different correlation and stacking programs, which utilize adaptive stacking and multi-channel cross-correlation, to get relative arrival time data for a set of high quality events. Selected earthquakes were larger than magnitude 5 and between 30 and 95 degrees away from the center of the array. Stations that consistently show late arrivals may have a low <span class="hlt">velocity</span> body beneath them, more clearly visualized via a 3-D tomographic model. Initial results from the two tomography methods indicate the presence of low-<span class="hlt">velocity</span> zones at several depths. Better resolved <span class="hlt">velocity</span> models will be developed as more data are acquired.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19850016756&hterms=Solar+power+filters&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DSolar%2Bpower%2Bfilters','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19850016756&hterms=Solar+power+filters&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DSolar%2Bpower%2Bfilters"><span>Evaluation of a Magneto-optical Filter and a Fabry-perot Interferometer for the Measurement of Solar <span class="hlt">Velocity</span> <span class="hlt">Fields</span> from Space</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rhodes, E. J., Jr.; Cacciani, A.; Blamont, J.; Tomczyk, S.; Ulrich, R. K.; Howard, R. F.</p> <p>1984-01-01</p> <p>A program was developed to evaluate the performance of three different devices as possible space-borne solar <span class="hlt">velocity</span> <span class="hlt">field</span> imagers. Two of these three devices, a magneto-optical filter and a molecular adherence Fabry-Perot interferometer were installed in a newly-constructed <span class="hlt">observing</span> system located at the 60-foot tower telescope at the Mt. Wilson Observatory. Time series of solar filtergrams and Dopplergrams lasting up to 10 hours per day were obtained with the filter while shorter runs were obtained with the Fabry-Perot. Two-dimensional k (sub h)-omega power spectra which show clearly the well-known p-mode ridges were computed from the time series obtained with the magneto-optical filter. These power spectra were compared with similar power spectra obtained recently with the 13.7-m McMath spectrograph at Kitt Peak.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880037707&hterms=sars&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsars','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880037707&hterms=sars&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsars"><span><span class="hlt">Observation</span> of sea-ice dynamics using synthetic aperture radar images: Automated analysis</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Vesecky, John F.; Samadani, Ramin; Smith, Martha P.; Daida, Jason M.; Bracewell, Ronald N.</p> <p>1988-01-01</p> <p>The European Space Agency's ERS-1 satellite, as well as others planned to follow, is expected to carry synthetic-aperture radars (SARs) over the polar regions beginning in 1989. A key component in utilization of these SAR data is an automated scheme for extracting the sea-ice <span class="hlt">velocity</span> <span class="hlt">field</span> from a time sequence of SAR images of the same geographical region. Two techniques for automated sea-ice tracking, image pyramid area correlation (hierarchical correlation) and feature tracking, are described. Each technique is applied to a pair of Seasat SAR sea-ice images. The results compare well with each other and with manually tracked estimates of the ice <span class="hlt">velocity</span>. The advantages and disadvantages of these automated methods are pointed out. Using these ice <span class="hlt">velocity</span> <span class="hlt">field</span> estimates it is possible to construct one sea-ice image from the other member of the pair. Comparing the reconstructed image with the <span class="hlt">observed</span> image, errors in the estimated <span class="hlt">velocity</span> <span class="hlt">field</span> can be recognized and a useful probable error display created automatically to accompany ice <span class="hlt">velocity</span> estimates. It is suggested that this error display may be useful in segmenting the sea ice <span class="hlt">observed</span> into regions that move as rigid plates of significant ice <span class="hlt">velocity</span> shear and distortion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1429109-observations-field-aligned-ion-ion-beam-instability-magnetized-laboratory-plasma','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1429109-observations-field-aligned-ion-ion-beam-instability-magnetized-laboratory-plasma"><span><span class="hlt">Observations</span> of a <span class="hlt">field</span>-aligned ion/ion-beam instability in a magnetized laboratory plasma</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Heuer, P. V.; Weidl, M. S.; Dorst, R. S.; ...</p> <p>2018-03-01</p> <p>Collisionless coupling between super Alfvénic ions and an ambient plasma parallel to a background magnetic <span class="hlt">field</span> is mediated by a set of electromagnetic ion/ion-beam instabilities including the resonant right hand instability (RHI). To study this coupling and its role in parallel shock formation, a new experimental configuration at the University of California, Los Angeles utilizes high-energy and high-repetition-rate lasers to create a super-Alfvénic <span class="hlt">field</span>-aligned debris plasma within an ambient plasma in the Large Plasma Device. We used a time-resolved fluorescence monochromator and an array of Langmuir probes to characterize the laser plasma <span class="hlt">velocity</span> distribution and density. The debris ions weremore » <span class="hlt">observed</span> to be sufficiently super-Alfvénic and dense to excite the RHI. Measurements with magnetic flux probes exhibited a right-hand circularly polarized frequency chirp consistent with the excitation of the RHI near the laser target. To conclude, we compared measurements to 2D hybrid simulations of the experiment.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1429109-observations-field-aligned-ion-ion-beam-instability-magnetized-laboratory-plasma','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1429109-observations-field-aligned-ion-ion-beam-instability-magnetized-laboratory-plasma"><span><span class="hlt">Observations</span> of a <span class="hlt">field</span>-aligned ion/ion-beam instability in a magnetized laboratory 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>Heuer, P. V.; Weidl, M. S.; Dorst, R. S.</p> <p></p> <p>Collisionless coupling between super Alfvénic ions and an ambient plasma parallel to a background magnetic <span class="hlt">field</span> is mediated by a set of electromagnetic ion/ion-beam instabilities including the resonant right hand instability (RHI). To study this coupling and its role in parallel shock formation, a new experimental configuration at the University of California, Los Angeles utilizes high-energy and high-repetition-rate lasers to create a super-Alfvénic <span class="hlt">field</span>-aligned debris plasma within an ambient plasma in the Large Plasma Device. We used a time-resolved fluorescence monochromator and an array of Langmuir probes to characterize the laser plasma <span class="hlt">velocity</span> distribution and density. The debris ions weremore » <span class="hlt">observed</span> to be sufficiently super-Alfvénic and dense to excite the RHI. Measurements with magnetic flux probes exhibited a right-hand circularly polarized frequency chirp consistent with the excitation of the RHI near the laser target. To conclude, we compared measurements to 2D hybrid simulations of the experiment.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910011708&hterms=mourning+duel+psychoanalysis&qs=Ntx%3Dmode%2Bmatchany%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dmourning%2Bduel%2Bpsychoanalysis','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910011708&hterms=mourning+duel+psychoanalysis&qs=Ntx%3Dmode%2Bmatchany%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dmourning%2Bduel%2Bpsychoanalysis"><span>Fabry-Perot <span class="hlt">observations</span> of comet Austin</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Schultz, David; Scherb, F.; Roesler, F. L.; Li, G.; Harlander, J.; Roberts, T. P. P.; Vandenberk, D.; Nossal, S.; Coakley, M.; Oliversen, Ronald J.</p> <p>1990-01-01</p> <p>Preliminary results of a program to <span class="hlt">observe</span> Comet Austin (1990c1) from 16 April to 4 May and from 11 May to 27 May 1990 using the West Auxiliary of the McMath Solar Telescope on Kitt Peak, Arizona were presetned. The <span class="hlt">observations</span> were made with a 15 cm duel-etalon Fabry-Perot scanning and imaging spectrometer with two modes of operation: a high resolution mode with a <span class="hlt">velocity</span> resolution of 1.2 km/s and a medium resolution mode with a <span class="hlt">velocity</span> resolution 10 km/s. Scanning data was obtained with an RCA C31034A photomultiplier tube and imaging data was obtained with a Photometrics LN2 cooled CCD camera with a 516 by 516 Ford chip. The results include: (1) information on the coma outflow <span class="hlt">velocity</span> from high resolution spectral profiles of (OI)6300 and NH2 emissions, (2) gaseous water production rates from medium resolution <span class="hlt">observation</span> of (OI)6300, (3) spectra of H2O(+) emissions in order to study the ionized component of the coma, (4) spatial distribution of H2O(+) emission features from sequences of <span class="hlt">velocity</span> resolved images (data cubes), and (5) spatial distribution of (OI)6300 and NH2 emissions from medium resolution images. The <span class="hlt">field</span> of view on the sky was 10.5 arcminutes in diameter. In the imaging mode the CCD was binned 4 by 4 resulting in 7.6 sec power pixel and a subarray readout for a <span class="hlt">field</span> of view of 10.5 min.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70025650','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70025650"><span>Gas transfer <span class="hlt">velocities</span> measured at low wind speed over a lake</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Crusius, John; Wanninkhof, R.</p> <p>2003-01-01</p> <p>The relationship between gas transfer <span class="hlt">velocity</span> and wind speed was evaluated at low wind speeds by quantifying the rate of evasion of the deliberate tracer, SF6, from a small oligotrophic lake. Several possible relationships between gas transfer <span class="hlt">velocity</span> and low wind speed were evaluated by using 1-min-averaged wind speeds as a measure of the instantaneous wind speed values. Gas transfer <span class="hlt">velocities</span> in this data set can be estimated virtually equally well by assuming any of three widely used relationships between k600 and winds referenced to 10-m height, U10: (1) a bilinear dependence with a break in the slope at ???3.7 m s-1, which resulted in the best fit; (2) a power dependence; and (3) a constant transfer <span class="hlt">velocity</span> for U10 3.7 m s-1 which, coupled with the typical variability in instantaneous wind speeds <span class="hlt">observed</span> in the <span class="hlt">field</span>, leads to average transfer <span class="hlt">velocity</span> estimates that are higher than those predicted for steady wind trends. The transfer <span class="hlt">velocities</span> predicted by the bilinear steady wind relationship for U10 < ???3.7 m s-1 are virtually identical to the theoretical predictions for transfer across a smooth surface.</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 <span class="hlt">Field</span> and Mission Deep <span class="hlt">Field</span> 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 gradients and overburden geopressure gradients 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/2015EGUGA..17.9771G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..17.9771G"><span>Modeling continuous seismic <span class="hlt">velocity</span> changes due to ground shaking in Chile</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gassenmeier, Martina; Richter, Tom; Sens-Schönfelder, Christoph; Korn, Michael; Tilmann, Frederik</p> <p>2015-04-01</p> <p>In order to investigate temporal seismic <span class="hlt">velocity</span> changes due to earthquake related processes and environmental forcing, we analyze 8 years of ambient seismic noise recorded by the Integrated Plate Boundary Observatory Chile (IPOC) network in northern Chile between 18° and 25° S. The Mw 7.7 Tocopilla earthquake in 2007 and the Mw 8.1 Iquique earthquake in 2014 as well as numerous smaller events occurred in this area. By autocorrelation of the ambient seismic noise <span class="hlt">field</span>, approximations of the Green's functions are retrieved. The recovered function represents backscattered or multiply scattered energy from the immediate neighborhood of the station. To detect relative changes of the seismic <span class="hlt">velocities</span> we apply the stretching method, which compares individual autocorrelation functions to stretched or compressed versions of a long term averaged reference autocorrelation function. We use time windows in the coda of the autocorrelations, that contain scattered waves which are highly sensitive to minute changes in the <span class="hlt">velocity</span>. At station PATCX we <span class="hlt">observe</span> seasonal changes in seismic <span class="hlt">velocity</span> as well as temporary <span class="hlt">velocity</span> reductions in the frequency range of 4-6 Hz. The seasonal changes can be attributed to thermal stress changes in the subsurface related to variations of the atmospheric temperature. This effect can be modeled well by a sine curve and is subtracted for further analysis of short term variations. Temporary <span class="hlt">velocity</span> reductions occur at the time of ground shaking usually caused by earthquakes and are followed by a recovery. 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. Our hypothesis is that not only the shaking of earthquakes provokes <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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008cosp...37..555C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008cosp...37..555C"><span>Electromagnetic pulse scattering by a wedge moving in a free space with relativistic <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>Ciarkowski, Adam</p> <p></p> <p>Recently, increased interest is <span class="hlt">observed</span> in studying scattering of electromagnetic signals by objects moving with large <span class="hlt">velocities</span>. The <span class="hlt">velocities</span> considered can attain relativistic values. Interesting phenomena characteristic of this class of problems were <span class="hlt">observed</span>, in this number the Doppler shift of equiphase surfaces in the diffracted wave. Apart from new techniques elaborated to attack general scattering problems involving moving objects, specific scaterring problems are also examined. Of special interest are moving scatterers with edges. The simplest scaterrer with this property is a wedge, which in particular case reduces to a half-plane. There is a number of recent works in which diffraction of specific electromagnetic signals by these objects in motion are analyzed. In most cases time-harmonic excitation <span class="hlt">fields</span> are being assumed. This contribution is concerned with the analysis of 2D scattering of an electromagnetic pulse by a perfectly conducting wedge moving in a free space with relativistic <span class="hlt">velocity</span>. The exciting <span class="hlt">field</span> is a pulsed plane-wave signal, with its envelope described by a Dirac delta function. This choice is motivated by the fact that solutions to excitation <span class="hlt">fields</span> with different envelopes can be obtained from that found here by its integration with an appropriate weight function. In this sense this solution plays a role of a Green function. In our analysis we neglect any dispersion phenomena connected with the surrounding medium. The results herein obtained may be useful in modelling phenomena connected with the space technology. In our analysis we apply the Frame Hopping Method. In particular we first Lorentz transform the pulse signal from the laboratory frame of reference where this <span class="hlt">field</span> is defined, to the frame where the wedge is at rest. In the latter frame we Fourier transform the resulting <span class="hlt">field</span> to the complex frequency domain, thus arriving at the problem of time-harmonic diffraction by the wedge at rest. This problem has the exact</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15246418','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15246418"><span><span class="hlt">Velocity</span> <span class="hlt">field</span> measurement in gas-liquid metal two-phase flow with use of PIV and neutron radiography techniques.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Saito, Y; Mishima, K; Tobita, Y; Suzuki, T; Matsubayashi, M</p> <p>2004-10-01</p> <p>To establish reasonable safety concepts for the realization of commercial liquid-metal fast breeder reactors, it is indispensable to demonstrate that the release of excessive energy due to re-criticality of molten core could be prevented even if a severe core damage accident took place. Two-phase flow due to the boiling of fuel-steel mixture in the molten core pool has a larger liquid-to-gas density ratio and higher surface tension in comparison with those of ordinary two-phase flows such as air-water flow. In this study, to investigate the effect of the recirculation flow on the bubble behavior, visualization and measurement of nitrogen gas-molten lead bismuth in a rectangular tank was performed by using neutron radiography and particle image velocimetry techniques. Measured flow parameters include flow regime, two-dimensional void distribution, and liquid <span class="hlt">velocity</span> <span class="hlt">field</span> in the tank. The present technique is applicable to the measurement of <span class="hlt">velocity</span> <span class="hlt">fields</span> and void fraction, and the basic characteristics of gas-liquid metal two-phase mixture were clarified.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5776005-magnetic-monopole-plasma-oscillations-survival-galactic-magnetic-fields','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5776005-magnetic-monopole-plasma-oscillations-survival-galactic-magnetic-fields"><span>Magnetic monopole plasma oscillations and the survival of Galactic magnetic <span class="hlt">fields</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>Parker, E.N.</p> <p></p> <p>This paper explores the general nature of magnetic-monopole plasma oscillations as a theoretical possibility for the <span class="hlt">observed</span> Galactic magnetic <span class="hlt">field</span> in the presence of a high abundance of magnetic monopoles. The modification of the hydromagnetic induction equation by the monopole oscillations produces the half-<span class="hlt">velocity</span> effect, in which the magnetic <span class="hlt">field</span> is transported bodily with a <span class="hlt">velocity</span> midway between the motion of the conducting fluid and the monopole plasma. <span class="hlt">Observational</span> studies of the magnetic <span class="hlt">field</span> in the Galaxy, and in other galaxies, exclude the half-<span class="hlt">velocity</span> effect, indicating that the magnetic <span class="hlt">fields</span> is not associated with monopole oscillations. In any case themore » phase mixing would destroy the oscillations in less than 100 Myr. The conclusion is that magnetic monopole oscillations do not play a significant role in the galactic magnetic <span class="hlt">fields</span>. Hence the existence of galactic magnetic <span class="hlt">fields</span> places a low limit on the monopole flux, so that their detection - if they exist at all - requires a collecting area at least as large as a football <span class="hlt">field</span>. 47 references.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSH21C2550V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSH21C2550V"><span>Study of <span class="hlt">Velocity</span> and Magnetic <span class="hlt">Field</span> Fluctuations at Kinetic Scale with the DSCOVR Data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vech, D.; Kasper, J. C.; Klein, K. G.; Hegedus, A. M.; Stevens, M. L.; Case, A. W.; Szabo, A.; Koval, A.</p> <p>2016-12-01</p> <p>The Deep Space Climate Observatory (DSCOVR), launched in 2015, performs high resolution measurements of the solar wind at the L1 vantage point. The Faraday cup onboard DSCOVR is capable of sampling solar wind <span class="hlt">velocity</span> distribution functions at cadences up to 1 Hz, which is complemented by the 50 samples/sec magnetic <span class="hlt">field</span> experiment. The combined usage of these data makes it possible to study kinetic scale physics, in particular turbulent fluctuations and the associated dissipation processes with unprecedented resolution. In this work we investigate recently obtained data sets and analyze correlations between the magnetic <span class="hlt">field</span> and the measured currents in different energy/charge windows. The goal of the study is to search for active wave-particle interactions at specific locations in phase space. We estimate the significance of these correlations and discuss the implications for our understanding of kinetic scale physics of the solar wind.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006GeoJI.166..991W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006GeoJI.166..991W"><span>Geodetic <span class="hlt">observations</span> of ice flow <span class="hlt">velocities</span> over the southern part of subglacial Lake Vostok, Antarctica, and their glaciological implications</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wendt, Jens; Dietrich, Reinhard; Fritsche, Mathias; Wendt, Anja; Yuskevich, Alexander; Kokhanov, Andrey; Senatorov, Anton; Lukin, Valery; Shibuya, Kazuo; Doi, Koichiro</p> <p>2006-09-01</p> <p>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 <span class="hlt">velocities</span> 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 <span class="hlt">velocity</span> 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 <span class="hlt">velocity</span> 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 <span class="hlt">velocities</span> 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 <span class="hlt">observed</span> geodetic quantities yields an accretion rate in the same order of magnitude. Furthermore, the comparison of our geodetic <span class="hlt">observations</span> with results inferred from ice-penetrating radar data indicates that the ice flow may not have changed significantly for several thousand years.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017CQGra..34w4001V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017CQGra..34w4001V"><span>Simulations of extragalactic magnetic <span class="hlt">fields</span> and of their <span class="hlt">observables</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vazza, F.; Brüggen, M.; Gheller, C.; Hackstein, S.; Wittor, D.; Hinz, P. M.</p> <p>2017-12-01</p> <p>The origin of extragalactic magnetic <span class="hlt">fields</span> is still poorly understood. Based on a dedicated suite of cosmological magneto-hydrodynamical simulations with the ENZO code we have performed a survey of different models that may have caused present-day magnetic <span class="hlt">fields</span> in galaxies and galaxy clusters. The outcomes of these models differ in cluster outskirts, filaments, sheets and voids and we use these simulations to find <span class="hlt">observational</span> signatures of magnetogenesis. With these simulations, we predict the signal of extragalactic magnetic <span class="hlt">fields</span> in radio <span class="hlt">observations</span> of synchrotron emission from the cosmic web, in Faraday rotation, in the propagation of ultra high energy cosmic rays, in the polarized signal from fast radio bursts at cosmological distance and in spectra of distant blazars. In general, primordial scenarios in which present-day magnetic <span class="hlt">fields</span> originate from the amplification of weak (⩽nG ) uniform seed <span class="hlt">fields</span> result in more homogeneous and relatively easier to <span class="hlt">observe</span> magnetic <span class="hlt">fields</span> than astrophysical scenarios, in which present-day <span class="hlt">fields</span> are the product of feedback processes triggered by stars and active galaxies. In the near future the best evidence for the origin of cosmic magnetic <span class="hlt">fields</span> will most likely come from a combination of synchrotron emission and Faraday rotation <span class="hlt">observed</span> at the periphery of large-scale structures.</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> </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('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('http://adsabs.harvard.edu/abs/2018ApJ...856..127G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...856..127G"><span>IRIS <span class="hlt">Observations</span> of Magnetic Interactions in the Solar Atmosphere between Preexisting and Emerging Magnetic <span class="hlt">Fields</span>. I. Overall Evolution</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Guglielmino, Salvo L.; Zuccarello, Francesca; Young, Peter R.; Murabito, Mariarita; Romano, Paolo</p> <p>2018-04-01</p> <p>We report multiwavelength ultraviolet <span class="hlt">observations</span> taken with the IRIS satellite, concerning the emergence phase in the upper chromosphere and transition region of an emerging flux region (EFR) embedded in the preexisting <span class="hlt">field</span> of active region NOAA 12529 in the Sun. IRIS data are complemented by full-disk <span class="hlt">observations</span> of the Solar Dynamics Observatory satellite, relevant to the photosphere and the corona. The photospheric configuration of the EFR is also analyzed by measurements taken with the spectropolarimeter on board the Hinode satellite, when the EFR was fully developed. Recurrent intense brightenings that resemble UV bursts, with counterparts in all coronal passbands, are identified at the edges of the EFR. Jet activity is also <span class="hlt">observed</span> at chromospheric and coronal levels, near the <span class="hlt">observed</span> brightenings. The analysis of the IRIS line profiles reveals the heating of dense plasma in the low solar atmosphere and the driving of bidirectional high-<span class="hlt">velocity</span> flows with speed up to 100 km s‑1 at the same locations. Compared with previous <span class="hlt">observations</span> and numerical models, these signatures suggest evidence of several long-lasting, small-scale magnetic reconnection episodes between the emerging bipole and the ambient <span class="hlt">field</span>. This process leads to the cancellation of a preexisting photospheric flux concentration and appears to occur higher in the atmosphere than usually found in UV bursts, explaining the <span class="hlt">observed</span> coronal counterparts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22521600-coronal-magnetic-fields-derived-from-simultaneous-microwave-euv-observations-comparison-potential-field-model','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22521600-coronal-magnetic-fields-derived-from-simultaneous-microwave-euv-observations-comparison-potential-field-model"><span>CORONAL MAGNETIC <span class="hlt">FIELDS</span> DERIVED FROM SIMULTANEOUS MICROWAVE AND EUV <span class="hlt">OBSERVATIONS</span> AND COMPARISON WITH THE POTENTIAL <span class="hlt">FIELD</span> MODEL</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>Miyawaki, Shun; Nozawa, Satoshi; Iwai, Kazumasa</p> <p>2016-02-10</p> <p>We estimated the accuracy of coronal magnetic <span class="hlt">fields</span> derived from radio <span class="hlt">observations</span> by comparing them to potential <span class="hlt">field</span> calculations and the differential emission measure measurements using EUV <span class="hlt">observations</span>. We derived line-of-sight components of the coronal magnetic <span class="hlt">field</span> from polarization <span class="hlt">observations</span> of the thermal bremsstrahlung in the NOAA active region 11150, <span class="hlt">observed</span> around 3:00 UT on 2011 February 3 using the Nobeyama Radioheliograph at 17 GHz. Because the thermal bremsstrahlung intensity at 17 GHz includes both chromospheric and coronal components, we extracted only the coronal component by measuring the coronal emission measure in EUV <span class="hlt">observations</span>. In addition, we derived only themore » radio polarization component of the corona by selecting the region of coronal loops and weak magnetic <span class="hlt">field</span> strength in the chromosphere along the line of sight. The upper limits of the coronal longitudinal magnetic <span class="hlt">fields</span> were determined as 100–210 G. We also calculated the coronal longitudinal magnetic <span class="hlt">fields</span> from the potential <span class="hlt">field</span> extrapolation using the photospheric magnetic <span class="hlt">field</span> obtained from the Helioseismic and Magnetic Imager. However, the calculated potential <span class="hlt">fields</span> were certainly smaller than the <span class="hlt">observed</span> coronal longitudinal magnetic <span class="hlt">field</span>. This discrepancy between the potential and the <span class="hlt">observed</span> magnetic <span class="hlt">field</span> strengths can be explained consistently by two reasons: (1) the underestimation of the coronal emission measure resulting from the limitation of the temperature range of the EUV <span class="hlt">observations</span>, and (2) the underestimation of the coronal magnetic <span class="hlt">field</span> resulting from the potential <span class="hlt">field</span> assumption.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeoJI.213...30I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeoJI.213...30I"><span>Monitoring of the spatio-temporal change in the interplate coupling at northeastern Japan subduction zone based on the spatial gradients of surface <span class="hlt">velocity</span> <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Iinuma, Takeshi</p> <p>2018-04-01</p> <p>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 <span class="hlt">fields</span> 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 <span class="hlt">observations</span>. The gradient of the surface <span class="hlt">velocities</span> 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 <span class="hlt">velocities</span> 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 <span class="hlt">velocity</span> 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 <span class="hlt">field</span> 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 <span class="hlt">field</span> 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 <span class="hlt">observation</span> 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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MNRAS.467.1386B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MNRAS.467.1386B"><span>Pairwise <span class="hlt">velocities</span> in the "Running FLRW" cosmological model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bibiano, Antonio; Croton, Darren J.</p> <p>2017-05-01</p> <p>We present an analysis of the pairwise <span class="hlt">velocity</span> statistics from a suite of cosmological N-body simulations describing the 'Running Friedmann-Lemaître-Robertson-Walker' (R-FLRW) cosmological model. This model is based on quantum <span class="hlt">field</span> theory in a curved space-time and extends Λ cold dark matter (CDM) with a time-evolving vacuum energy density, ρ _Λ. To enforce local conservation of matter, a time-evolving gravitational coupling is also included. Our results constitute the first study of <span class="hlt">velocities</span> in the R-FLRW cosmology, and we also compare with other dark energy simulations suites, repeating the same analysis. We find a strong degeneracy between the pairwise <span class="hlt">velocity</span> and σ8 at z = 0 for almost all scenarios considered, which remains even when we look back to epochs as early as z = 2. We also investigate various coupled dark energy models, some of which show minimal degeneracy, and reveal interesting deviations from ΛCDM that could be readily exploited by future cosmological <span class="hlt">observations</span> to test and further constrain our understanding of dark energy.</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 <span class="hlt">field</span>. 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 <span class="hlt">field</span>, pressure gradient 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('http://adsabs.harvard.edu/abs/2005DSRII..52.1639M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005DSRII..52.1639M"><span>Comparisons between <span class="hlt">observations</span> and numerical simulations of Japan (East) Sea flow and mass <span class="hlt">fields</span> in 1999 through 2001</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mooers, Christopher N. K.; Bang, Inkweon; Sandoval, Francisco J.</p> <p>2005-06-01</p> <p>The Princeton Ocean Model (POM), as implemented for the Japan (East) Sea (JES) with mesoscale-admitting resolution is driven by seasonal throughflow and synoptic atmospheric forcing for 1999 through 2001. Temperature and salinity profiles from shipborne and PALACE float CTDs, and horizontal <span class="hlt">velocities</span> at 800 m from PALACE float trajectories, plus horizontal <span class="hlt">velocities</span> at 15 m from WOCE surface drifters for 1988 through 2001, are used to assess the performance of the numerical simulations for a base case. General agreement exists in the circulation at 15 and 800 m and the horizontal and vertical structure of the upper ocean temperature and salinity <span class="hlt">fields</span>. The mean <span class="hlt">observed</span> flow at 15 m defines the two branches of the Tsushima Warm Current and hints at the existence of a large cyclonic gyre over the Japan Basin, which the simulations also produce. The mean <span class="hlt">observed</span> flow at 800 m defines a large cyclonic recirculation gyre over the Japan Basin that validates the simulated flow pattern. Variances of the <span class="hlt">observed</span> and simulated flows at 15 and 800 m have similar patterns. The main discrepancies are associated with the strength of the seasonal thermocline and halocline and the location of the Subpolar Front. When smoother topography and smaller lateral friction are used in other cases, the thermocline and halocline strengthen, agreeing better with the <span class="hlt">observed</span> values, and when 80% of total outflow transport is forced to exit through Soya Strait, the Subpolar Front extends along the coast to the north of Tsugaru Strait, which is an <span class="hlt">observed</span> feature absent in the base case.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeoRL..45.3043H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeoRL..45.3043H"><span>Crustal Deformation across the Jericho Valley Section of the Dead Sea Fault as Resolved by Detailed <span class="hlt">Field</span> and Geodetic <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>Hamiel, Yariv; Piatibratova, Oksana; Mizrahi, Yaakov; Nahmias, Yoav; Sagy, Amir</p> <p>2018-04-01</p> <p>Detailed <span class="hlt">field</span> and geodetic <span class="hlt">observations</span> of crustal deformation across the Jericho Fault section of the Dead Sea Fault are presented. New <span class="hlt">field</span> <span class="hlt">observations</span> reveal several slip episodes that rupture the surface, consist with strike slip and extensional deformation along a fault zone width of about 200 m. Using dense Global Positioning System measurements, we obtain the <span class="hlt">velocities</span> of new stations across the fault. We find that this section is locked for strike-slip motion with a locking depth of 16.6 ± 7.8 km and a slip rate of 4.8 ± 0.7 mm/year. The Global Positioning System measurements also indicate asymmetrical extension at shallow depths of the Jericho Fault section, between 0.3 and 3 km. Finally, our results suggest the vast majority of the sinistral slip along the Dead Sea Fault in southern Jorden Valley is accommodated by the Jericho Fault section.</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('https://www.osti.gov/servlets/purl/1455317','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1455317"><span><span class="hlt">Field</span> measurement of <span class="hlt">velocity</span> time series in the center of Sequim Bay</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>Harding, Samuel F.; Harker-Klimes, Genevra EL</p> <p></p> <p>A 600 kHz RDI Workhorse was installed in the center of Sequim Bay from 15:04 June 23, 2017 to 09:34 August 24, 2017 at a depth of 25.9 m from MLLW. The instrument was configured to record the flow <span class="hlt">velocity</span> in vertical cells of 1.0 m in 10 minute ensembles. Each ensemble was calculated as the mean of 24 pings, sampled with an interval of 5.0 s. A burst of increased sampling rate (1200 samples at 2 Hz) was recorded to characterize the wave climate on an hourly basis. The peak depth-averaged flow speed for the deployment was recorded duringmore » the flood tide on June 24, 2017 with a magnitude of 0.34 m/s. The peak flow speed in a single bin was recorded during the same tide at a location of 11.6 m from the seabed with a magnitude of 0.46 m/s. The <span class="hlt">velocity</span> direction was <span class="hlt">observed</span> to be relatively constant as a function of depth for the higher flow <span class="hlt">velocities</span> (flood tides) but highly variable during times of slower flow (ebb tides). A peak significant wave height of 0.36 m was recorded on June 30, 2017 at 18:54. The measured waves showed no indication of a prevalent wave direction during this deployment. The wave record of the fetch-limited site during this deployment approaches the lower limit of the wave measurement resolution. The water temperature fluctuated over a range of 1.7°C during the deployment duration. The mean pitch of the instrument was -1.2° and the mean roll angle of the instrument was 0.3°. The low pitch and roll angles are important factors in the accurate measurement of the wave activity at the surface.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012EGUGA..14.6690G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012EGUGA..14.6690G"><span>Reservoir Changes Derived from Seismic <span class="hlt">Observations</span> at The Geysers Geothermal <span class="hlt">Field</span>, CA, USA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gritto, R.; Jarpre, S.</p> <p>2012-04-01</p> <p>Induced seismicity associated with the exploitation of geothermal <span class="hlt">fields</span> is used as a tool to characterize and delineate changes associated with injection and production of fluids from the reservoir. At the same time public concern of felt seismicity has led to objections against the operation of geothermal reservoirs in close proximity to population centers. Production at the EGS sites in Basel (Switzerland) was stopped after renewed seismicity caused concern and objection from the public in the city. Operations in other geothermal reservoirs had to be scaled back or interrupted due to an unexpected increase in seismicity (Soultz-sous-forêt, France, Berlín, El Salvador). As a consequence of these concerns and in order to optimize the use of induced seismicity for reservoir engineering purposes, it becomes imperative to understand the relationship between seismic events and stress changes in the reservoir. We will address seismicity trends at The Geysers Geothermal Reservoir, CA USA, to understand the role of historical seismicity associated with past injection of water and/or production of steam. Our analysis makes use of a comprehensive database of earthquakes and associated phase arrivals from 2004 to 2011. A high-precision sub-set of the earthquake data was selected to analyze temporal changes in seismic <span class="hlt">velocities</span> and Vp/Vs-ratio throughout the whole reservoir. We find relatively low Vp/Vs values in 2004 suggestive of a vapor dominated reservoir. With passing time, however, the <span class="hlt">observed</span> temporal increase in Vp/Vs, coupled with a decrease in P- and S-wave <span class="hlt">velocities</span> suggests the presence of fluid-filled fractured rock. Considering the start of a continuous water injection project in 2004, it can be concluded that the fluid saturation of the reservoir has successfully recovered. Preliminary results of 3-D <span class="hlt">velocity</span> inversions of seismic data appear to corroborate earlier findings that the lowest Vp/Vs estimates are <span class="hlt">observed</span> in the center of the reservoir</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.S51C2436W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.S51C2436W"><span>The wireless networking system of Earthquake precursor mobile <span class="hlt">field</span> <span class="hlt">observation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, C.; Teng, Y.; Wang, X.; Fan, X.; Wang, X.</p> <p>2012-12-01</p> <p>The mobile <span class="hlt">field</span> <span class="hlt">observation</span> network could be real-time, reliably record and transmit large amounts of data, strengthen the physical signal <span class="hlt">observations</span> in specific regions and specific period, it can improve the monitoring capacity and abnormal tracking capability. According to the features of scatter everywhere, a large number of current earthquake precursor <span class="hlt">observation</span> measuring points, networking technology is based on wireless broadband accessing McWILL system, the communication system of earthquake precursor mobile <span class="hlt">field</span> <span class="hlt">observation</span> would real-time, reliably transmit large amounts of data to the monitoring center from measuring points through the connection about equipment and wireless accessing system, broadband wireless access system and precursor mobile <span class="hlt">observation</span> management center system, thereby implementing remote instrument monitoring and data transmition. At present, the earthquake precursor <span class="hlt">field</span> mobile <span class="hlt">observation</span> network technology has been applied to fluxgate magnetometer array geomagnetic <span class="hlt">observations</span> of Tianzhu, Xichang,and Xinjiang, it can be real-time monitoring the working status of the <span class="hlt">observational</span> instruments of large area laid after the last two or three years, large scale <span class="hlt">field</span> operation. Therefore, it can get geomagnetic <span class="hlt">field</span> data of the local refinement regions and provide high-quality <span class="hlt">observational</span> data for impending earthquake tracking forecast. Although, wireless networking technology is very suitable for mobile <span class="hlt">field</span> <span class="hlt">observation</span> with the features of simple, flexible networking etc, it also has the phenomenon of packet loss etc when transmitting a large number of <span class="hlt">observational</span> data due to the wireless relatively weak signal and narrow bandwidth. In view of high sampling rate instruments, this project uses data compression and effectively solves the problem of data transmission packet loss; Control commands, status data and <span class="hlt">observational</span> data transmission use different priorities and means, which control the packet loss rate within</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23005411','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23005411"><span>Energy transport <span class="hlt">velocity</span> in bidispersed magnetic colloids.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bhatt, Hem; Patel, Rajesh; Mehta, R V</p> <p>2012-07-01</p> <p>Study of energy transport <span class="hlt">velocity</span> of light is an effective background for slow, fast, and diffuse light and exhibits the photonic property of the material. We report a theoretical analysis of magnetic <span class="hlt">field</span> dependent resonant behavior in forward-backward anisotropy factor, light diffusion constant, and energy transport <span class="hlt">velocity</span> for bidispersed magnetic colloids. A bidispersed magnetic colloid is composed of micrometer size magnetic spheres dispersed in a magnetic nanofluid consisting of magnetic nanoparticles in a nonmagnetic liquid carrier. Magnetic Mie resonances and reduction in energy transport <span class="hlt">velocity</span> accounts for the possible delay (longer dwell time) by <span class="hlt">field</span> dependent resonant light transport. This resonant behavior of light in bidispersed magnetic colloids suggests a novel magnetophotonic material.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19980235520','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19980235520"><span>Normal Component of Induced <span class="hlt">Velocity</span> for Entire <span class="hlt">Field</span> of a Uniformly Loaded Lifting Rotor with Highly Swept Wake as Determined by Electromagnetic Analog</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Castles, Walter, Jr.; Durham, Howard L., Jr.; Kevorkian, Jirair</p> <p>1959-01-01</p> <p>Values of the normal component of induced <span class="hlt">velocity</span> throughout the entire <span class="hlt">field</span> of a uniformly loaded r(rotor at high high speed are presented in the form of charts and tables. Many points were found by an electromagnetic analog, details of which are given. Comparisons of computed and analog values for the induced <span class="hlt">velocity</span> indicate that the latter are sufficiently accurate for engineering purposes.</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> <span class="hlt">field</span> of 7 Zw 403 shows no clear large-scale organized motion but the <span class="hlt">velocity</span> <span class="hlt">field</span> 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> <span class="hlt">field</span> 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> gradients 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/22547286','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22547286"><span><span class="hlt">Velocity</span> measurement by vibro-acoustic Doppler.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nabavizadeh, Alireza; Urban, Matthew W; Kinnick, Randall R; Fatemi, Mostafa</p> <p>2012-04-01</p> <p>We describe the theoretical principles of a new Doppler method, which uses the acoustic response of a moving object to a highly localized dynamic radiation force of the ultrasound <span class="hlt">field</span> to calculate the <span class="hlt">velocity</span> of the moving object according to Doppler frequency shift. This method, named vibro-acoustic Doppler (VAD), employs two ultrasound beams separated by a slight frequency difference, Δf, transmitting in an X-focal configuration. Both ultrasound beams experience a frequency shift because of the moving objects and their interaction at the joint focal zone produces an acoustic frequency shift occurring around the low-frequency (Δf) acoustic emission signal. The acoustic emission <span class="hlt">field</span> resulting from the vibration of the moving object is detected and used to calculate its <span class="hlt">velocity</span>. We report the formula that describes the relation between Doppler frequency shift of the emitted acoustic <span class="hlt">field</span> and the <span class="hlt">velocity</span> of the moving object. To verify the theory, we used a string phantom. We also tested our method by measuring fluid <span class="hlt">velocity</span> in a tube. The results show that the error calculated for both string and fluid <span class="hlt">velocities</span> is less than 9.1%. Our theory shows that in the worst case, the error is 0.54% for a 25° angle variation for the VAD method compared with an error of -82.6% for a 25° angle variation for a conventional continuous wave Doppler method. An advantage of this method is that, unlike conventional Doppler, it is not sensitive to angles between the ultrasound beams and direction of motion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70157554','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70157554"><span>Using micro-seismicity and seismic <span class="hlt">velocities</span> to map subsurface geologic and hydrologic structure within the Coso geothermal <span class="hlt">field</span>, 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>Kaven, Joern Ole; Hickman, Stephen H.; Davatzes, Nicholas C.</p> <p>2012-01-01</p> <p>Geothermal reservoirs derive their capacity for fluid and heat transport in large part from faults and fractures. Micro-seismicity generated on such faults and fractures can be used to map larger fault structures as well as secondary fractures that add access to hot rock, fluid storage and recharge capacity necessary to have a sustainable geothermal resource. Additionally, inversion of seismic <span class="hlt">velocities</span> from micro-seismicity permits imaging of regions subject to the combined effects of fracture density, fluid pressure and steam content, among other factors. We relocate 14 years of seismicity (1996-2009) in the Coso geothermal <span class="hlt">field</span> using differential travel times and simultaneously invert for seismic <span class="hlt">velocities</span> to improve our knowledge of the subsurface geologic and hydrologic structure. We utilize over 60,000 micro-seismic events using waveform cross-correlation to augment to expansive catalog of P- and S-wave differential travel times recorded at Coso. We further carry out rigorous uncertainty estimation and find that our results are precise to within 10s of meters of relative location error. We find that relocated micro-seismicity outlines prominent, through-going faults in the reservoir in some cases. We also find that a significant portion of seismicity remains diffuse and does not cluster into more sharply defined major structures. The seismic <span class="hlt">velocity</span> structure reveals heterogeneous distributions of compressional (Vp) and shear (Vs) wave speed, with Vp generally lower in the main <span class="hlt">field</span> when compared to the east flank and Vs varying more significantly in the shallow portions of the reservoir. The Vp/Vs ratio appears to outline the two main compartments of the reservoir at depths of -0.5 to 1.5 km (relative to sea-level), with a ridge of relatively high Vp/Vs separating the main <span class="hlt">field</span> from the east flank. In the deeper portion of the reservoir this ridge is less prominent. Our results indicate that high-precision relocations of micro-seismicity can provide</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70025413','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70025413"><span>Anisotropic changes in P-wave <span class="hlt">velocity</span> and attenuation during deformation and fluid infiltration of granite</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Stanchits, S.A.; Lockner, D.A.; Ponomarev, A.V.</p> <p>2003-01-01</p> <p>Fluid infiltration and pore fluid pressure changes are known to have a significant effect on the occurrence of earthquakes. Yet, for most damaging earthquakes, with nucleation zones below a few kilometers depth, direct measurements of fluid pressure variations are not available. Instead, pore fluid pressures are inferred primarily from seismic-wave propagation characteristics such as Vp/Vs ratio, attenuation, and reflectivity contacts. We present laboratory measurements of changes in P-wave <span class="hlt">velocity</span> and attenuation during the injection of water into a granite sample as it was loaded to failure. A cylindrical sample of Westerly granite was deformed at constant confining and pore pressures of 50 and 1 MPa, respectively. Axial load was increased in discrete steps by controlling axial displacement. Anisotropic P-wave <span class="hlt">velocity</span> and attenuation <span class="hlt">fields</span> were determined during the experiment using an array of 13 piezoelectric transducers. At the final loading steps (86% and 95% of peak stress), both spatial and temporal changes in P-wave <span class="hlt">velocity</span> and peak-to-peak amplitudes of P and S waves were <span class="hlt">observed</span>. P-wave <span class="hlt">velocity</span> anisotropy reached a maximum of 26%. Transient increases in attenuation of up to 483 dB/m were also <span class="hlt">observed</span> and were associated with diffusion of water into the sample. We show that <span class="hlt">velocity</span> and attenuation of P waves are sensitive to the process of opening of microcracks and the subsequent resaturation of these cracks as water diffuses in from the surrounding region. Symmetry of the orientation of newly formed microcracks results in anisotropic <span class="hlt">velocity</span> and attenuation <span class="hlt">fields</span> that systematically evolve in response to changes in stress and influx of water. With proper scaling, these measurements provide constraints on the magnitude and duration of <span class="hlt">velocity</span> and attenuation transients that can be expected to accompany the nucleation of earthquakes in the Earth's crust.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJ...856...21C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...856...21C"><span>Diagnosing the Magnetic <span class="hlt">Field</span> Structure of a Coronal Cavity <span class="hlt">Observed</span> during the 2017 Total Solar Eclipse</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Yajie; Tian, Hui; Su, Yingna; Qu, Zhongquan; Deng, Linhua; Jibben, Patricia R.; Yang, Zihao; Zhang, Jingwen; Samanta, Tanmoy; He, Jiansen; Wang, Linghua; Zhu, Yingjie; Zhong, Yue; Liang, Yu</p> <p>2018-03-01</p> <p>We present an investigation of a coronal cavity <span class="hlt">observed</span> above the western limb in the coronal red line Fe X 6374 Å using a telescope of Peking University and in the green line Fe XIV 5303 Å using a telescope of Yunnan Observatories, Chinese Academy of Sciences, during the total solar eclipse on 2017 August 21. A series of magnetic <span class="hlt">field</span> models is constructed based on the magnetograms taken by the Helioseismic and Magnetic Imager on board the Solar Dynamics Observatory (SDO) one week before the eclipse. The model <span class="hlt">field</span> lines are then compared with coronal structures seen in images taken by the Atmospheric Imaging Assembly on board SDO and in our coronal red line images. The best-fit model consists of a flux rope with a twist angle of 3.1π, which is consistent with the most probable value of the total twist angle of interplanetary flux ropes <span class="hlt">observed</span> at 1 au. Linear polarization of the Fe XIII 10747 Å line calculated from this model shows a “lagomorphic” signature that is also <span class="hlt">observed</span> by the Coronal Multichannel Polarimeter of the High Altitude Observatory. We also find a ring-shaped structure in the line-of-sight <span class="hlt">velocity</span> of Fe XIII 10747 Å, which implies hot plasma flows along a helical magnetic <span class="hlt">field</span> structure, in the cavity. These results suggest that the magnetic structure of the cavity is a highly twisted flux rope, which may erupt eventually. The temperature structure of the cavity has also been investigated using the intensity ratio of Fe XIII 10747 Å and Fe X 6374 Å.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017IAUS..327...34J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017IAUS..327...34J"><span>Granular cells in the presence of magnetic <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jurčák, J.; Lemmerer, B.; van Noort, M.</p> <p>2017-10-01</p> <p>We present a statistical study of the dependencies of the shapes and sizes of the photospheric convective cells on the magnetic <span class="hlt">field</span> properties. This analysis is based on a 2.5 hour long SST <span class="hlt">observations</span> of active region NOAA 11768. We have blue continuum images taken with a cadence of 5.6 sec that are used for segmentation of individual granules and 270 maps of spectropolarimetric CRISP data allowing us to determine the properties of the magnetic <span class="hlt">field</span> along with the line-of-sight <span class="hlt">velocities</span>. The sizes and shapes of the granular cells are dependent on the the magnetic <span class="hlt">field</span> strength, where the granules tend to be smaller in regions with stronger magnetic <span class="hlt">field</span>. In the presence of highly inclined magnetic <span class="hlt">fields</span>, the eccentricity of granules is high and we do not <span class="hlt">observe</span> symmetric granules in these regions. The mean up-flow <span class="hlt">velocities</span> in granules as well as the granules intensities decrease with increasing magnetic <span class="hlt">field</span> strength.</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('https://ntrs.nasa.gov/search.jsp?R=19860037048&hterms=Electric+current&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DElectric%2Bcurrent','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19860037048&hterms=Electric+current&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DElectric%2Bcurrent"><span><span class="hlt">Observations</span> of <span class="hlt">field</span>-aligned currents, waves, and electric <span class="hlt">fields</span> at substorm onset</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Smits, D. P.; Hughes, W. J.; Cattell, C. A.; Russell, C. T.</p> <p>1986-01-01</p> <p>Substorm onsets, identified Pi 2 pulsations <span class="hlt">observed</span> on the Air Force Geophysics Laboratory Magnetometer Network, are studied using magnetometer and electric <span class="hlt">field</span> data from ISEE 1 as well as magnetometer data from the geosynchronous satellites GOES 2 and 3. The mid-latitude magnetometer data provides the means of both timing and locating the substorm onset so that the spacecraft locations with respect to the substorm current systems are known. During two intervals, each containing several onsets or intensifications, ISEE 1 <span class="hlt">observed</span> <span class="hlt">field</span>-aligned current signatures beginning simultaneously with the mid-latitude Pi 2 pulsation. Close to the earth broadband bursts of wave noise were <span class="hlt">observed</span> in the electric <span class="hlt">field</span> data whenever <span class="hlt">field</span>-aligned currents were detected. One onset occurred when ISEE 1 and GOES 2 were on the same <span class="hlt">field</span> line but in opposite hemispheres. During this onset ISEE 1 and GOES 2 saw magnetic signatures which appear to be due to conjugate <span class="hlt">field</span>-aligned currents flowing out of the western end of the westward auroral electrojets. The ISEE 1 signature is of a line current moving westward past the spacecraft. During the other interval, ISEE 1 was in the near-tail region near the midnight meridian. Plasma data confirms that the plasma sheet thinned and subsequently expanded at onset. Electric <span class="hlt">field</span> data shows that the plasma moved in the opposite direction to the plasma sheet boundary as the boundary expanded which implies that there must have been an abundant source of hot plasma present. The plasma motion was towards the center of the plasma sheet and earthwards and consisted of a series of pulses rather than a steady flow.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170003582&hterms=electric&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Delectric','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170003582&hterms=electric&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Delectric"><span>MMS Multipoint Electric <span class="hlt">Field</span> <span class="hlt">Observations</span> of Small-Scale Magnetic Holes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Goodrich, Katherine A.; Ergun, Robert E.; Wilder, Frederick; Burch, James; Torbert, Roy; Khotyaintsev, Yuri; Lindqvist, Per-Arne; Russell, Christopher; Strangeway, Robert; Magnus, Werner</p> <p>2016-01-01</p> <p>Small-scale magnetic holes (MHs), local depletions in magnetic <span class="hlt">field</span> strength, have been <span class="hlt">observed</span> multiple times in the Earths magnetosphere in the bursty bulk flow (BBF) braking region. This particular subset of MHs has <span class="hlt">observed</span> scale sizes perpendicular to the background magnetic <span class="hlt">field</span> (B) less than the ambient ion Larmor radius (p(sib i)). Previous <span class="hlt">observations</span> by Time History of Events and Macroscale Interactions during Substorms (THEMIS) indicate that this subset of MHs can be supported by a current driven by the E x B drift of electrons. Ions do not participate in the E x B drift due to the small-scale size of the electric <span class="hlt">field</span>. While in the BBF braking region, during its commissioning phase, the Magnetospheric Multiscale (MMS) spacecraft <span class="hlt">observed</span> a small-scale MH. The electric <span class="hlt">field</span> <span class="hlt">observations</span> taken during this event suggest the presence of electron currents perpendicular to the magnetic <span class="hlt">field</span>. These <span class="hlt">observations</span> also suggest that these currents can evolve to smaller spatial scales.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018IAUS..333...83P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018IAUS..333...83P"><span>MWA <span class="hlt">Observations</span> of the EOR1 <span class="hlt">Field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pindor, Bart</p> <p>2018-05-01</p> <p>The EOR1 <span class="hlt">field</span> at (RA,DEC)=(4hrs, -30.0°) is one of the main targets of the MWA EOR experiment. It is notable for being in one of the coldest regions of the southern radio sky, as well as for containing the bright radio galaxy Fornax A. We report an early demonstration that the distance of this <span class="hlt">field</span> from the Galactic Centre may make it a prime <span class="hlt">field</span> for EOR <span class="hlt">observations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1436640-situ-measurement-velocity-stress-sensitivity-using-crosswell-continuous-active-source-seismic-monitoring','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1436640-situ-measurement-velocity-stress-sensitivity-using-crosswell-continuous-active-source-seismic-monitoring"><span>In situ measurement of <span class="hlt">velocity</span>-stress sensitivity using crosswell continuous active-source seismic monitoring</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>Marchesini, P; Ajo-Franklin, JB; Daley, TM</p> <p>2017-09-01</p> <p>© 2017 Society of Exploration Geophysicists. The ability to characterize time-varying reservoir properties, such as the state of stress, has fundamental implications in subsurface engineering, relevant to geologic sequestration of CO2. Stress variation, here in the form of changes in pore fluid pressure, is one factor known to affect seismic <span class="hlt">velocity</span>. Induced variations in <span class="hlt">velocity</span> have been used in seismic studies to determine and monitor changes in the stress state. Previous studies conducted to determine <span class="hlt">velocity</span>-stress sensitivity at reservoir conditions rely primarily on laboratory measurements of core samples or theoretical relationships. We have developed a novel <span class="hlt">field</span>-scale experiment designed tomore » study the in situ relationship between pore-fluid pressure and seismic <span class="hlt">velocity</span> using a crosswell continuous active-source seismic monitoring (CASSM) system. At the Cranfield, Mississippi, CO2 sequestration <span class="hlt">field</span> site, we actively monitored seismic response for five days with a temporal resolution of 5 min; the target was a 26 m thick injection zone at approximately 3.2 km depth in a fluvial sandstone formation (lower Tuscaloosa Formation). The variation of pore fluid pressure was obtained during discrete events of fluid withdrawal from one of the two wells and monitored with downhole pressure sensors. The results indicate a correlation between decreasing CASSM time delay (i.e., <span class="hlt">velocity</span> change for a raypath in the reservoir) and periods of reduced fluid pore pressure. The correlation is interpreted as the <span class="hlt">velocity</span>-stress sensitivity measured in the reservoir. This <span class="hlt">observation</span> is consistent with published laboratory studies documenting a <span class="hlt">velocity</span> (V) increase with an effective stress increase. A traveltime change (dt) of 0.036 ms is measured as the consequence of a change in pressure of approximately 2.55 MPa (dPe). For T 1/4 13 ms total traveltime, the <span class="hlt">velocity</span>-stress sensitivity is dV/V/dPe 1/4 dt/T/dPe 1/4 10.9 × 10-4/MPa. The overall results suggest</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPC.1857d0006S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1857d0006S"><span>Rotation and strain rate of Sulawesi from geometrical <span class="hlt">velocity</span> <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sarsito, D. A.; Susilo, Simons, W. J. F.; Abidin, H. Z.; Sapiie, B.; Triyoso, W.; Andreas, H.</p> <p>2017-07-01</p> <p>One of methods that can be used to determine the tectonic deformation status is rate estimation from geometric rotation and strain using quantitative <span class="hlt">velocity</span> data from GPS <span class="hlt">observations</span>. Microplate Sulawesi region located in the zone of triple junction (Eurasia, Australia and Philippine Sea Plates) has very complex tectonic and seismic condition, which is why become very important to know its recent deformation status in order to study neo-tectonic and disaster mitigation. Deformation rate quantification is estimated in two parameters: rotation and geodetic strain rate of each GPS station Delaunay triangle in the study area. The analysis in this study is not done using the grids since there is no rheological information at location that can be used as the interpolation-extrapolation constraints. Our analysis reveals that Sulawesi is characterized by rapid rotation in several different domains and compression-strain pattern that varies depending on the type and boundary conditions of microplate. This information is useful for studying neo tectonic deformation status and earthquake disaster mitigation.</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> <span class="hlt">Fields</span> 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 <span class="hlt">fields</span> 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> gradients inside the liquid metal influence the propagation of ultrasonic waves. The effect of temperature and <span class="hlt">velocity</span> gradients 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 gradients representative for MYRRHA conditions. Once validated in water, the same model is used to make predictions for the effect of gradients in the MYRRHA liquid metal environment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRC..122.3519A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRC..122.3519A"><span>Downstream evolution of the Kuroshio's time-varying transport and <span class="hlt">velocity</span> structure</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Andres, M.; Mensah, V.; Jan, S.; Chang, M.-H.; Yang, Y.-J.; Lee, C. M.; Ma, B.; Sanford, T. B.</p> <p>2017-05-01</p> <p><span class="hlt">Observations</span> from two companion <span class="hlt">field</span> programs—Origins of the Kuroshio and Mindanao Current (OKMC) and <span class="hlt">Observations</span> of Kuroshio Transport Variability (OKTV)—are used here to examine the Kuroshio's temporal and spatial evolution. Kuroshio strength and <span class="hlt">velocity</span> structure were measured between June 2012 and November 2014 with pressure-sensor equipped inverted echo sounders (PIESs) and upward-looking acoustic Doppler current profilers (ADCPs) deployed across the current northeast of Luzon, Philippines, and east of Taiwan with an 8 month overlap in the two arrays' deployment periods. The time-mean net (i.e., integrated from the surface to the bottom) absolute transport increases downstream from 7.3 Sv (±4.4 Sv standard error) northeast of Luzon to 13.7 Sv (±3.6 Sv) east of Taiwan. The <span class="hlt">observed</span> downstream increase is consistent with the return flow predicted by the simple Sverdrup relation and the mean wind stress curl <span class="hlt">field</span> over the North Pacific (despite the complicated bathymetry and gaps along the North Pacific western boundary). Northeast of Luzon, the Kuroshio—bounded by the 0 m s-1 isotach—is shallower than 750 dbar, while east of Taiwan areas of positive flow reach to the seafloor (3000 m). Both arrays indicate a deep counterflow beneath the poleward-flowing Kuroshio (-10.3 ± 2.3 Sv by Luzon and -12.5 ± 1.2 Sv east of Taiwan). Time-varying transports and <span class="hlt">velocities</span> indicate the strong influence at both sections of westward propagating eddies from the ocean interior. Topography associated with the ridges east of Taiwan also influences the mean and time-varying <span class="hlt">velocity</span> structure there.</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> <span class="hlt">field</span> 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> gradients, 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://adsabs.harvard.edu/abs/2014yCat..51480118N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014yCat..51480118N"><span>VizieR Online Data Catalog: Radial <span class="hlt">velocity</span> curves of LMC ellipsoidal variables (Nie+, 2014)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nie, J. D.; Wood, P. R.</p> <p>2014-11-01</p> <p>We initially selected 86 sequence E candidates from those given in Soszynski et al. 2004 (cat. J/AcA/54/347). The radial <span class="hlt">velocity</span> <span class="hlt">observations</span> were taken using the Wide <span class="hlt">Field</span> Spectrograph (WiFeS) mounted on the Australian National University 2.3m telescope at Siding Spring Observatory. WiFes has six gratings. For our <span class="hlt">observations</span>, the gratings B7000 (wavelength coverage of 4184-5580Å) and I7000 (wavelength coverage of 6832-9120Å) were chosen for the blue and red CCD, respectively. These two gratings give a two-pixel resolution R=7000. We carried out 18 weeks of radial <span class="hlt">velocity</span> monitoring, from 2010 September to 2012 March. (2 data files).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SPIE.9827E..0FM','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SPIE.9827E..0FM"><span><span class="hlt">Velocity</span> <span class="hlt">fields</span> and optical turbulence near the boundary in a strongly convective laboratory flow</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Matt, Silvia; Hou, Weilin; Goode, Wesley; Hellman, Samuel</p> <p>2016-05-01</p> <p>Boundary layers around moving underwater vehicles or other platforms can be a limiting factor for optical communication. Turbulence in the boundary layer of a body moving through a stratified medium can lead to small variations in the index of refraction, which impede optical signals. As a first step towards investigating this boundary layer effect on underwater optics, we study the flow near the boundary in the Rayleigh-Bénard laboratory tank at the Naval Research Laboratory Stennis Space Center. The tank is set up to generate temperature-driven, i.e., convective turbulence, and allows control of the turbulence intensity. This controlled turbulence environment is complemented by computational fluid dynamics simulations to visualize and quantify multi-scale flow patterns. The boundary layer dynamics in the laboratory tank are quantified using a state-of-the-art Particle Image Velocimetry (PIV) system to examine the boundary layer <span class="hlt">velocities</span> and turbulence parameters. The <span class="hlt">velocity</span> <span class="hlt">fields</span> and flow dynamics from the PIV are compared to the numerical model and show the model to accurately reproduce the <span class="hlt">velocity</span> range and flow dynamics. The temperature variations and thus optical turbulence effects can then be inferred from the model temperature data. Optical turbulence is also visible in the raw data from the PIV system. The newly collected data are consistent with previously reported measurements from high-resolution Acoustic Doppler Velocimeter profilers (Nortek Vectrino), as well as fast thermistor probes and novel next-generation fiber-optics temperature sensors. This multi-level approach to studying optical turbulence near a boundary, combining in-situ measurements, optical techniques, and numerical simulations, can provide new insight and aid in mitigating turbulence impacts on underwater optical signal transmission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015ApJ...800...83B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015ApJ...800...83B"><span>The Power Spectrum of the Milky Way: <span class="hlt">Velocity</span> Fluctuations in the Galactic Disk</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bovy, Jo; Bird, Jonathan C.; García Pérez, Ana E.; Majewski, Steven R.; Nidever, David L.; Zasowski, Gail</p> <p>2015-02-01</p> <p>We investigate the kinematics of stars in the mid-plane of the Milky Way (MW) on scales between 25 pc and 10 kpc with data from the Apache Point Observatory Galactic Evolution Experiment (APOGEE), the Radial <span class="hlt">Velocity</span> Experiment (RAVE), and the Geneva-Copenhagen survey (GCS). Using red-clump (RC) stars in APOGEE, we determine the large-scale line-of-sight <span class="hlt">velocity</span> <span class="hlt">field</span> out to 5 kpc from the Sun in (0.75 kpc)2 bins. The solar motion V ⊙ - c with respect to the circular <span class="hlt">velocity</span> Vc is the largest contribution to the power on large scales after subtracting an axisymmetric rotation <span class="hlt">field</span>; we determine the solar motion by minimizing the large-scale power to be V ⊙ - c = 24 ± 1 (ran.) ± 2 (syst. [Vc ]) ± 5 (syst.[large-scale]) km s-1, where the systematic uncertainty is due to (1) a conservative 20 km s-1 uncertainty in Vc and (2) the estimated power on unobserved larger scales. Combining the APOGEE peculiar-<span class="hlt">velocity</span> <span class="hlt">field</span> with RC stars in RAVE out to 2 kpc from the Sun and with local GCS stars, we determine the power spectrum of residual <span class="hlt">velocity</span> fluctuations in the MW's disk on scales between 0.2 kpc-1 <= k <= 40 kpc-1. Most of the power is contained in a broad peak between 0.2 kpc-1 < k < 0.9 kpc-1. We investigate the expected power spectrum for various non-axisymmetric perturbations and demonstrate that the central bar with commonly used parameters but of relatively high mass can explain the bulk of <span class="hlt">velocity</span> fluctuations in the plane of the Galactic disk near the Sun. Streaming motions ≈10 km s-1 on >~ 3 kpc scales in the MW are in good agreement with <span class="hlt">observations</span> of external galaxies and directly explain why local determinations of the solar motion are inconsistent with global measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20080045749&hterms=VR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DVR','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20080045749&hterms=VR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DVR"><span>Cassini <span class="hlt">Observations</span> of Saturn's Magnetotail Region: Preliminary Results</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sittler, E. C.; Arridge, C.; Rymer, A.; Coates, A.; Krupp, N.; Blanc, M.; Richardson, J.; Andre, N.; Thomsen, M.; Tokar, R. L.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20080045749'); toggleEditAbsImage('author_20080045749_show'); toggleEditAbsImage('author_20080045749_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20080045749_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20080045749_hide"></p> <p>2007-01-01</p> <p>Using Cassini thermal plasma, hot plasma and magnetic <span class="hlt">field</span> <span class="hlt">observations</span> for several intervals between the dawn meridian of Saturn's outer magnetosphere and Saturn's magnetotail region, we investigate the structure of the magnetotail, plasma and magnetic <span class="hlt">field</span> properties within tail-like current sheet regions and ion flows within the magnetotail regions. We use Cassini Plasma Spectrometer (CAPS) Ion Mass Spectrometer (IMS), Electron Plasma Spectrometer (ELS) <span class="hlt">observations</span>, MIMI LEMMS ion and electron <span class="hlt">observations</span> and Cassini magnetometer data (MAG) to characterize the plasma environment. IMS <span class="hlt">observations</span> are used to measure plasma flow <span class="hlt">velocities</span> from which one can infer rotation versus convective flows. IMS composition measurements are used to trace the source of plasma from the inner magnetosphere (protons, H2+ and water group ions) versus an external solar wind source (protons and e +i+on s). A critical parameter for both models is the strength of the convection electric <span class="hlt">field</span> with respect to the rotational electric <span class="hlt">field</span> for the large scale magnetosphere. For example, are there significant return flows (i.e., negative radial <span class="hlt">velocities</span>, VR < 0) and/or plasmoids (V(sub R) > 0) within the magnetotail region? Initial preliminary evidence of such out flows and return flows was presented by Sittler et al. This talk complements the more global analysis by McAndrews et al.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFM.G41C..04N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFM.G41C..04N"><span>New GPS <span class="hlt">velocity</span> <span class="hlt">field</span> in the northern Andes (Peru - Ecuador - Colombia): heterogeneous locking along the subduction, northeastwards motion of the Northern Andes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nocquet, J.; Mothes, P. A.; Villegas Lanza, J.; Chlieh, M.; Jarrin, P.; Vallée, M.; Tavera, H.; Ruiz, G.; Regnier, M.; Rolandone, F.</p> <p>2010-12-01</p> <p>Rapid subduction of the Nazca plate beneath the northen Andes margin (~6 cm/yr) results in two different processes: (1) elastic stress is accumulating along the Nazca/South American plate interface which is responsible for one of the largest megathrust earthquake sequences during the last century. The 500-km-long rupture zone of the 1906 (Mw= 8.8) event was partially reactivated by three events from the 1942 (Mw = 7.8), 1958 (Mw = 7.7), to the 1979 (Mw = 8.2). However, south of latitude 1°S, no M>8 earthquake has been reported in the last three centuries, suggesting that this area is slipping aseismically (2) permanent deformation causes opening of the Gulf of Guayaquil, with northeastwards motion of the Northern Andean Block (NAB). We present a new GPS <span class="hlt">velocity</span> <span class="hlt">field</span> covering the northern Andes from south of the Gulf of Guayaquil to the Caribbean plate. Our <span class="hlt">velocity</span> <span class="hlt">field</span> includes new continuously-recording GPS stations installed along the Ecuadorian coast, together with campaign sites <span class="hlt">observed</span> since 1994 in the CASA project (Kellogg et al., 1989). We first estimate the long-term kinematics of the NAB in a joint inversion including GPS data, earthquake slip vectors, and quaternary slip rates on major faults. The inversion provides an Euler pole located at long. -107.8°E, lat. 36.2°N, 0.091°/Ma and indicates little internal deformation of the NAB (wrms=1.2 mm/yr). As a consequence, 30% of the obliquity of the Nazca/South America motion is accommodated by transcurrent to transpressive motion along the eastern boundary of the NAB. Residual <span class="hlt">velocities</span> with respect to the NAB are then modeled in terms Models indicate a patchwork of highly coupled asperities encompassed by aseismic patches over the area of rupture of the M~8.8 1906 earthquake. Very low coupling is found along the southern Ecuadorian and northern Peru subduction.</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 <span class="hlt">field</span> 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 gradient from Baade's Window (l,b) = 0.9°, -3.9° to our innermost <span class="hlt">field</span>, in contrast to the marked gradient <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://pubs.er.usgs.gov/publication/70033413','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70033413"><span>A study of methods to estimate debris flow <span class="hlt">velocity</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Prochaska, A.B.; Santi, P.M.; Higgins, J.D.; Cannon, S.H.</p> <p>2008-01-01</p> <p>Debris flow <span class="hlt">velocities</span> are commonly back-calculated from superelevation events which require subjective estimates of radii of curvature of bends in the debris flow channel or predicted using flow equations that require the selection of appropriate rheological models and material property inputs. This research investigated difficulties associated with the use of these conventional <span class="hlt">velocity</span> estimation methods. Radii of curvature estimates were found to vary with the extent of the channel investigated and with the scale of the media used, and back-calculated <span class="hlt">velocities</span> varied among different investigated locations along a channel. Distinct populations of Bingham properties were found to exist between those measured by laboratory tests and those back-calculated from <span class="hlt">field</span> data; thus, laboratory-obtained values would not be representative of <span class="hlt">field</span>-scale debris flow behavior. To avoid these difficulties with conventional methods, a new preliminary <span class="hlt">velocity</span> estimation method is presented that statistically relates flow <span class="hlt">velocity</span> to the channel slope and the flow depth. This method presents ranges of reasonable <span class="hlt">velocity</span> predictions based on 30 previously measured <span class="hlt">velocities</span>. ?? 2008 Springer-Verlag.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvB..97h5302S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvB..97h5302S"><span>Optimization of edge state <span class="hlt">velocity</span> in the integer quantum Hall regime</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sahasrabudhe, H.; Novakovic, B.; Nakamura, J.; Fallahi, S.; Povolotskyi, M.; Klimeck, G.; Rahman, R.; Manfra, M. J.</p> <p>2018-02-01</p> <p><span class="hlt">Observation</span> of interference in the quantum Hall regime may be hampered by a small edge state <span class="hlt">velocity</span> due to finite phase coherence time. Therefore designing two quantum point contact (QPCs) interferometers having a high edge state <span class="hlt">velocity</span> is desirable. Here we present a new simulation method for designing heterostructures with high edge state <span class="hlt">velocity</span> by realistically modeling edge states near QPCs in the integer quantum Hall effect (IQHE) regime. Using this simulation method, we also predict the filling factor at the center of QPCs and their conductance at different gate voltages. The 3D Schrödinger equation is split into 1D and 2D parts. Quasi-1D Schrödinger and Poisson equations are solved self-consistently in the IQHE regime to obtain the potential profile, and quantum transport is used to solve for the edge state wave functions. The <span class="hlt">velocity</span> of edge states is found to be <E >/B , where <E > is the expectation value of the electric <span class="hlt">field</span> for the edge state. Anisotropically etched trench gated heterostructures with double-sided delta doping have the highest edge state <span class="hlt">velocity</span> among the structures considered.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27677897','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27677897"><span>Impaired <span class="hlt">Velocity</span> Processing Reveals an Agnosia for Motion in Depth.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Barendregt, Martijn; Dumoulin, Serge O; Rokers, Bas</p> <p>2016-11-01</p> <p>Many individuals with normal visual acuity are unable to discriminate the direction of 3-D motion in a portion of their visual <span class="hlt">field</span>, a deficit previously referred to as a stereomotion scotoma. The origin of this visual deficit has remained unclear. We hypothesized that the impairment is due to a failure in the processing of one of the two binocular cues to motion in depth: changes in binocular disparity over time or interocular <span class="hlt">velocity</span> differences. We isolated the contributions of these two cues and found that sensitivity to interocular <span class="hlt">velocity</span> differences, but not changes in binocular disparity, varied systematically with <span class="hlt">observers</span>' ability to judge motion direction. We therefore conclude that the inability to interpret motion in depth is due to a failure in the neural mechanisms that combine <span class="hlt">velocity</span> signals from the two eyes. Given these results, we argue that the deficit should be considered a prevalent but previously unrecognized agnosia specific to the perception of visual motion. © The Author(s) 2016.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSM33E..05B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSM33E..05B"><span>Effects of the reconnection electric <span class="hlt">field</span> on crescent electron distribution functions in asymmetric guide <span class="hlt">field</span> reconnection</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bessho, N.; Chen, L. J.; Hesse, M.; Wang, S.</p> <p>2017-12-01</p> <p> <span class="hlt">velocity</span> space. We compare theory and PIC simulation results of the <span class="hlt">velocity</span> shift of crescent distribution functions based on the derived time period of bounce motion in a guide <span class="hlt">field</span>. Theoretical predictions are applied to electron distributions <span class="hlt">observed</span> by MMS in magnetopause reconnection to estimate the reconnection electric <span class="hlt">field</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24462149','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24462149"><span>In vivo lateral blood flow <span class="hlt">velocity</span> measurement using speckle size estimation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xu, Tiantian; Hozan, Mohsen; Bashford, Gregory R</p> <p>2014-05-01</p> <p>In previous studies, we proposed blood measurement using speckle size estimation, which estimates the lateral component of blood flow within a single image frame based on the <span class="hlt">observation</span> that the speckle pattern corresponding to blood reflectors (typically red blood cells) stretches (i.e., is "smeared") if blood flow is in the same direction as the electronically controlled transducer line selection in a 2-D image. In this <span class="hlt">observational</span> study, the clinical viability of ultrasound blood flow <span class="hlt">velocity</span> measurement using speckle size estimation was investigated and compared with that of conventional spectral Doppler of carotid artery blood flow data collected from human patients in vivo. Ten patients (six male, four female) were recruited. Right carotid artery blood flow data were collected in an interleaved fashion (alternating Doppler and B-mode A-lines) with an Antares Ultrasound Imaging System and transferred to a PC via the Axius Ultrasound Research Interface. The scanning <span class="hlt">velocity</span> was 77 cm/s, and a 4-s interval of flow data were collected from each subject to cover three to five complete cardiac cycles. Conventional spectral Doppler data were collected simultaneously to compare with estimates made by speckle size estimation. The results indicate that the peak systolic <span class="hlt">velocities</span> measured with the two methods are comparable (within ±10%) if the scan <span class="hlt">velocity</span> is greater than or equal to the flow <span class="hlt">velocity</span>. When scan <span class="hlt">velocity</span> is slower than peak systolic <span class="hlt">velocity</span>, the speckle stretch method asymptotes to the scan <span class="hlt">velocity</span>. Thus, the speckle stretch method is able to accurately measure pure lateral flow, which conventional Doppler cannot do. In addition, an initial comparison of the speckle size estimation and color Doppler methods with respect to computational complexity and data acquisition time indicated potential time savings in blood flow <span class="hlt">velocity</span> estimation using speckle size estimation. Further studies are needed for calculation of the speckle stretch method</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JVGR..356..114L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JVGR..356..114L"><span>A generic model for the shallow <span class="hlt">velocity</span> structure of volcanoes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lesage, Philippe; Heap, Michael J.; Kushnir, Alexandra</p> <p>2018-05-01</p> <p>The knowledge of the structure of volcanoes and of the physical properties of volcanic rocks is of paramount importance to the understanding of volcanic processes and the interpretation of monitoring <span class="hlt">observations</span>. However, the determination of these structures by geophysical methods suffers limitations including a lack of resolution and poor precision. Laboratory experiments provide complementary information on the physical properties of volcanic materials and their behavior as a function of several parameters including pressure and temperature. Nevertheless combined studies and comparisons of <span class="hlt">field</span>-based geophysical and laboratory-based physical approaches remain scant in the literature. Here, we present a meta-analysis which compares 44 seismic <span class="hlt">velocity</span> models of the shallow structure of eleven volcanoes, laboratory <span class="hlt">velocity</span> measurements on about one hundred rock samples from five volcanoes, and seismic well-logs from deep boreholes at two volcanoes. The comparison of these measurements confirms the strong variability of P- and S-wave <span class="hlt">velocities</span>, which reflects the diversity of volcanic materials. The values obtained from laboratory experiments are systematically larger than those provided by seismic models. This discrepancy mainly results from scaling problems due to the difference between the sampled volumes. The averages of the seismic models are characterized by very low <span class="hlt">velocities</span> at the surface and a strong <span class="hlt">velocity</span> increase at shallow depth. By adjusting analytical functions to these averages, we define a generic model that can describe the variations in P- and S-wave <span class="hlt">velocities</span> in the first 500 m of andesitic and basaltic volcanoes. This model can be used for volcanoes where no structural information is available. The model can also account for site time correction in hypocenter determination as well as for site and path effects that are commonly <span class="hlt">observed</span> in volcanic structures.</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('http://adsabs.harvard.edu/abs/2017A%26A...602A..75R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017A%26A...602A..75R"><span><span class="hlt">Observations</span> of apparent superslow wave propagation in solar prominences</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Raes, J. O.; Van Doorsselaere, T.; Baes, M.; Wright, A. N.</p> <p>2017-06-01</p> <p>Context. Phase mixing of standing continuum Alfvén waves and/or continuum slow waves in atmospheric magnetic structures such as coronal arcades can create the apparent effect of a wave propagating across the magnetic <span class="hlt">field</span>. Aims: We <span class="hlt">observe</span> a prominence with SDO/AIA on 2015 March 15 and find the presence of oscillatory motion. We aim to demonstrate that interpreting this motion as a magneto hydrodynamic (MHD) wave is faulty. We also connect the decrease of the apparent <span class="hlt">velocity</span> over time with the phase mixing process, which depends on the curvature of the magnetic <span class="hlt">field</span> lines. Methods: By measuring the displacement of the prominence at different heights to calculate the apparent <span class="hlt">velocity</span>, we show that the propagation slows down over time, in accordance with the theoretical work of Kaneko et al. We also show that this propagation speed drops below what is to be expected for even slow MHD waves for those circumstances. We use a modified Kippenhahn-Schlüter prominence model to calculate the curvature of the magnetic <span class="hlt">field</span> and fit our <span class="hlt">observations</span> accordingly. Results: Measuring three of the apparent waves, we get apparent <span class="hlt">velocities</span> of 14, 8, and 4 km s-1. Fitting a simple model for the magnetic <span class="hlt">field</span> configuration, we obtain that the filament is located 103 Mm below the magnetic centre. We also obtain that the scale of the magnetic <span class="hlt">field</span> strength in the vertical direction plays no role in the concept of apparent superslow waves and that the moment of excitation of the waves happened roughly one oscillation period before the end of the eruption that excited the oscillation. Conclusions: Some of the <span class="hlt">observed</span> phase <span class="hlt">velocities</span> are lower than expected for slow modes for the circumstances, showing that they rather fit with the concept of apparent superslow propagation. A fit with our magnetic <span class="hlt">field</span> model allows for inferring the magnetic geometry of the prominence. The movie attached to Fig. 1 is available at http://www.aanda.org</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28758031','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28758031"><span>Comparison of Retinal Microvessel Blood Flow <span class="hlt">Velocities</span> Acquired with Two Different <span class="hlt">Fields</span> of View.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhou, Jin; Li, Min; Chen, Wan; Yang, Ye; Hu, Liang; Wang, Liang; Jiang, Hong; Wang, Jianhua</p> <p>2017-01-01</p> <p>To compare the different retinal blood flow <span class="hlt">velocities</span> (BFVs) acquired with different <span class="hlt">fields</span> of view (FOVs) using the retinal function imager (RFI), twenty eyes of twenty healthy subjects were enrolled in the study. Retinal microvessel BFV in the macula was acquired with both a wide FOV (35 degrees, 7.3 × 7.3 mm 2 ) and a commonly used small FOV (20 degrees, 4.3 × 4.3 mm 2 ). The 35-degree FOV was trimmed to be equivalent to the 20-degree FOV to compare the BFVs of the similar FOVs using different settings. With the 35-degree FOV, both retinal arteriolar and venular BFVs were significantly greater than the 20-degree FOV ( P < 0.001). When the 20-degree FOV was compared to the trimmed equivalent 20-degree FOV acquired using the 35-degree FOV, significant BFV differences were found in both the arterioles ( P = 0.029) and venules ( P < 0.001). This is the first study to compare retinal blood flow <span class="hlt">velocities</span> acquired with different FOVs using RFI. The conversion factor from 35 degrees to 20 degrees is 0.95 for arteriolar BFV and 0.92 for venular BFV, which may be used for comparing BFVs acquired with different FOVs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5516753','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5516753"><span>Comparison of Retinal Microvessel Blood Flow <span class="hlt">Velocities</span> Acquired with Two Different <span class="hlt">Fields</span> of View</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Zhou, Jin; Chen, Wan; Yang, Ye; Hu, Liang; Wang, Liang</p> <p>2017-01-01</p> <p>To compare the different retinal blood flow <span class="hlt">velocities</span> (BFVs) acquired with different <span class="hlt">fields</span> of view (FOVs) using the retinal function imager (RFI), twenty eyes of twenty healthy subjects were enrolled in the study. Retinal microvessel BFV in the macula was acquired with both a wide FOV (35 degrees, 7.3 × 7.3 mm2) and a commonly used small FOV (20 degrees, 4.3 × 4.3 mm2). The 35-degree FOV was trimmed to be equivalent to the 20-degree FOV to compare the BFVs of the similar FOVs using different settings. With the 35-degree FOV, both retinal arteriolar and venular BFVs were significantly greater than the 20-degree FOV (P < 0.001). When the 20-degree FOV was compared to the trimmed equivalent 20-degree FOV acquired using the 35-degree FOV, significant BFV differences were found in both the arterioles (P = 0.029) and venules (P < 0.001). This is the first study to compare retinal blood flow <span class="hlt">velocities</span> acquired with different FOVs using RFI. The conversion factor from 35 degrees to 20 degrees is 0.95 for arteriolar BFV and 0.92 for venular BFV, which may be used for comparing BFVs acquired with different FOVs. PMID:28758031</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://adsabs.harvard.edu/abs/1993MNRAS.263..481V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1993MNRAS.263..481V"><span>Voids in Gravitational Instability Scenarios - Part One - Global Density and <span class="hlt">Velocity</span> <span class="hlt">Fields</span> in an Einstein - De-Sitter Universe</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 Weygaert, R.; van Kampen, E.</p> <p>1993-07-01</p> <p>The first results of an extensive study of the structure and dynamics of underdense regions in gravitational instability scenarios are presented. Instead of adopting spherically symmetric voids with some idealized initial density and <span class="hlt">velocity</span> profile, underdense regions of a given size and depth, embedded in an initial density fluctuation <span class="hlt">field</span>, are generated. In order to accomplish this in a consistent way, these initial conditions are set up by means of Bertschinger's constrained random <span class="hlt">field</span> code. The generated particle samples of 64^3^ particles in a box of side 100 Mpc are followed into the non-linear regime by Bertschinger's PM N- body code. In this way we address the dependence of the structure and kinematics of the void both on the initial depth of the void and on the fluctuation <span class="hlt">field</span> in which it is embedded. In particular, this study provides some understanding of how far fluctuations on small scales modify the dynamics of the large-scale void, and especially of how far the properties of small structures inside the void are affected by the global properties of the void. One of the conspicuous features of the initial density <span class="hlt">fields</span> inside protovoids appears to be the existence of a `void hierarchy', with small voids embedded in larger voids. The survival of this hierarchy during the riot evolution of the void depends critically on the initial depth as well as on the clustering scenario involved. As well as presenting a qualitative discussion of the structure of underdense regions in initial density <span class="hlt">fields</span> in different scenarios, and the results of simulations of the ensuing non-linear evolution, we concentrate in particular on a comparison of the global density and <span class="hlt">velocity</span> <span class="hlt">fields</span> in voids with predictions from linear theory as well as from the spherical outflow model. The relation between the initial linear depth, the resulting non-linear depth and the excess expansion <span class="hlt">velocities</span> in voids is addressed. In addition, we find that, while near its centre a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22130430-observation-low-magnetic-field-density-peaks-helicon-plasma','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22130430-observation-low-magnetic-field-density-peaks-helicon-plasma"><span><span class="hlt">Observation</span> of low magnetic <span class="hlt">field</span> density peaks in helicon 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>Barada, Kshitish K.; Chattopadhyay, P. K.; Ghosh, J.</p> <p>2013-04-15</p> <p>Single density peak has been commonly <span class="hlt">observed</span> in low magnetic <span class="hlt">field</span> (<100 G) helicon discharges. In this paper, we report the <span class="hlt">observations</span> of multiple density peaks in low magnetic <span class="hlt">field</span> (<100 G) helicon discharges produced in the linear helicon plasma device [Barada et al., Rev. Sci. Instrum. 83, 063501 (2012)]. Experiments are carried out using argon gas with m = +1 right helical antenna operating at 13.56 MHz by varying the magnetic <span class="hlt">field</span> from 0 G to 100 G. The plasma density varies with varying the magnetic <span class="hlt">field</span> at constant input power and gas pressure and reaches to its peakmore » value at a magnetic <span class="hlt">field</span> value of {approx}25 G. Another peak of smaller magnitude in density has been <span class="hlt">observed</span> near 50 G. Measurement of amplitude and phase of the axial component of the wave using magnetic probes for two magnetic <span class="hlt">field</span> values corresponding to the <span class="hlt">observed</span> density peaks indicated the existence of radial modes. Measured parallel wave number together with the estimated perpendicular wave number suggests oblique mode propagation of helicon waves along the resonance cone boundary for these magnetic <span class="hlt">field</span> values. Further, the <span class="hlt">observations</span> of larger floating potential fluctuations measured with Langmuir probes at those magnetic <span class="hlt">field</span> values indicate that near resonance cone boundary; these electrostatic fluctuations take energy from helicon wave and dump power to the plasma causing density peaks.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19920033132&hterms=deming&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Ddeming','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19920033132&hterms=deming&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Ddeming"><span>Absolute wind <span class="hlt">velocities</span> in the lower thermosphere of Venus using infrared heterodyne spectroscopy</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Goldstein, Jeffrey J.; Mumma, Michael J.; Kostiuk, Theodor; Deming, Drake; Espenak, Fred; Zipoy, David</p> <p>1991-01-01</p> <p>NASA's IR Telescope Facility and the McMath Solar Telescope have yielded absolute wind <span class="hlt">velocities</span> in the Venus thermosphere for December 1985 to March 1987 with sufficient spatial resolution for circulation model discrimination. A qualitative analysis of beam-integrated winds indicates subsolar-to-antisolar circulation in the lower thermosphere; horizontal wind <span class="hlt">velocity</span> was derived from a two-parameter model wind <span class="hlt">field</span> of subsolar-antisolar and zonal components. A unique model fit common to all <span class="hlt">observing</span> periods possessed 120 m/sec subsolar-antisolar and 25 m/sec zonal retrograde components, consistent with the Bougher et al. (1986, 1988) hydrodynamical models for 110 km.</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=19880033068&hterms=keefe&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D50%26Ntt%3Dkeefe','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880033068&hterms=keefe&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D50%26Ntt%3Dkeefe"><span>The size distributions of fragments ejected at a given <span class="hlt">velocity</span> from impact craters</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>O'Keefe, John D.; Ahrens, Thomas J.</p> <p>1987-01-01</p> <p>The mass distribution of fragments that are ejected at a given <span class="hlt">velocity</span> for impact craters is modeled to allow extrapolation of laboratory, <span class="hlt">field</span>, and numerical results to large scale planetary events. The model is semi-empirical in nature and is derived from: (1) numerical calculations of cratering and the resultant mass versus ejection <span class="hlt">velocity</span>, (2) <span class="hlt">observed</span> ejecta blanket particle size distributions, (3) an empirical relationship between maximum ejecta fragment size and crater diameter, (4) measurements and theory of maximum ejecta size versus ejecta <span class="hlt">velocity</span>, and (5) an assumption on the functional form for the distribution of fragments ejected at a given <span class="hlt">velocity</span>. This model implies that for planetary impacts into competent rock, the distribution of fragments ejected at a given <span class="hlt">velocity</span> is broad, e.g., 68 percent of the mass of the ejecta at a given <span class="hlt">velocity</span> contains fragments having a mass less than 0.1 times a mass of the largest fragment moving at that <span class="hlt">velocity</span>. The broad distribution suggests that in impact processes, additional comminution of ejecta occurs after the upward initial shock has passed in the process of the ejecta <span class="hlt">velocity</span> vector rotating from an initially downward orientation. This additional comminution produces the broader size distribution in impact ejecta as compared to that obtained in simple brittle failure experiments.</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> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22370577-very-large-array-green-bank-telescope-observations-orion-ngc-w12-photodissociation-region-properties-magnetic-field','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22370577-very-large-array-green-bank-telescope-observations-orion-ngc-w12-photodissociation-region-properties-magnetic-field"><span>Very large array and green bank telescope <span class="hlt">observations</span> of Orion B (NGC 2024, W12): photodissociation region properties and magnetic <span class="hlt">field</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>Roshi, D. Anish; Goss, W. M.; Jeyakumar, S., E-mail: aroshi@nrao.edu, E-mail: mgoss@nrao.edu, E-mail: sjk@astro.ugto.mx</p> <p></p> <p>We present images of C110α and H110α radio recombination line (RRL) emission at 4.8 GHz and images of H166α, C166α, and X166α RRL emission at 1.4 GHz, <span class="hlt">observed</span> toward the star-forming region NGC 2024. The 1.4 GHz image with angular resolution ∼70'' is obtained using Very Large Array (VLA) data. The 4.8 GHz image with angular resolution ∼17'' is obtained by combining VLA and Green Bank Telescope data in order to add the short and zero spacing data in the uv plane. These images reveal that the spatial distributions of C110α line emission is confined to the southern rim ofmore » the H II region close to the ionization front whereas the C166α line emission is extended in the north-south direction across the H II region. The LSR <span class="hlt">velocity</span> of the C110α line is 10.3 km s{sup –1} similar to that of lines <span class="hlt">observed</span> from molecular material located at the far side of the H II region. This similarity suggests that the photodissociation region (PDR) responsible for C110α line emission is at the far side of the H II region. The LSR <span class="hlt">velocity</span> of C166α is 8.8 km s{sup –1}. This <span class="hlt">velocity</span> is comparable with the <span class="hlt">velocity</span> of molecular absorption lines <span class="hlt">observed</span> from the foreground gas, suggesting that the PDR is at the near side of the H II region. Non-LTE models for carbon line-forming regions are presented. Typical properties of the foreground PDR are T {sub PDR} ∼ 100 K, n{sub e}{sup PDR}∼5 cm{sup –3}, n {sub H} ∼ 1.7 × 10{sup 4} cm{sup –3}, and path length l ∼ 0.06 pc, and those of the far side PDR are T {sub PDR} ∼ 200 K, n{sub e}{sup PDR}∼ 50 cm{sup –3}, n {sub H} ∼ 1.7 × 10{sup 5} cm{sup –3}, and l ∼ 0.03 pc. Our modeling indicates that the far side PDR is located within the H II region. We estimate the magnetic <span class="hlt">field</span> strength in the foreground PDR to be 60 μG and that in the far side PDR to be 220 μG. Our <span class="hlt">field</span> estimates compare well with the values obtained from OH Zeeman <span class="hlt">observations</span> toward NGC 2024. The H166α spectrum shows</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22365423-alignments-galaxies-around-virgo-cluster-local-velocity-shear','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22365423-alignments-galaxies-around-virgo-cluster-local-velocity-shear"><span>Alignments of the galaxies in and around the Virgo cluster with the local <span class="hlt">velocity</span> shear</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>Lee, Jounghun; Rey, Soo Chang; Kim, Suk, E-mail: jounghun@astro.snu.ac.kr</p> <p>2014-08-10</p> <p><span class="hlt">Observational</span> evidence is presented for the alignment between the cosmic sheet and the principal axis of the <span class="hlt">velocity</span> shear <span class="hlt">field</span> at the position of the Virgo cluster. The galaxies in and around the Virgo cluster from the Extended Virgo Cluster Catalog that was recently constructed by Kim et al. are used to determine the direction of the local sheet. The peculiar <span class="hlt">velocity</span> <span class="hlt">field</span> reconstructed from the Sloan Digital Sky Survey Data Release 7 is analyzed to estimate the local <span class="hlt">velocity</span> shear tensor at the Virgo center. Showing first that the minor principal axis of the local <span class="hlt">velocity</span> shear tensor ismore » almost parallel to the direction of the line of sight, we detect a clear signal of alignment between the positions of the Virgo satellites and the intermediate principal axis of the local <span class="hlt">velocity</span> shear projected onto the plane of the sky. Furthermore, the dwarf satellites are found to appear more strongly aligned than their normal counterparts, which is interpreted as an indication of the following. (1) The normal satellites and the dwarf satellites fall in the Virgo cluster preferentially along the local filament and the local sheet, respectively. (2) The local filament is aligned with the minor principal axis of the local <span class="hlt">velocity</span> shear while the local sheet is parallel to the plane spanned by the minor and intermediate principal axes. Our result is consistent with the recent numerical claim that the <span class="hlt">velocity</span> shear is a good tracer of the cosmic web.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009SPD....40.0916H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009SPD....40.0916H"><span>Solar Mean Magnetic <span class="hlt">Field</span> <span class="hlt">Observed</span> by GONG</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Harvey, J. W.; Petrie, G.; Clark, R.; GONG Team</p> <p>2009-05-01</p> <p>The average line-of-sight (LOS) magnetic <span class="hlt">field</span> of the Sun has been <span class="hlt">observed</span> for decades, either by measuring the circular polarization across a selected spectrum line using integrated sunlight or by averaging such measurements in spatially resolved images. The GONG instruments produce full-disk LOS magnetic images every minute, which can be averaged to yield the mean magnetic <span class="hlt">field</span> nearly continuously. Such measurements are well correlated with the heliospheric magnetic <span class="hlt">field</span> <span class="hlt">observed</span> near Earth about 4 days later. They are also a measure of solar activity on long and short time scales. Averaging a GONG magnetogram, with nominal noise of 3 G per pixel, results in a noise level of about 4 mG. This is low enough that flare-related <span class="hlt">field</span> changes have been seen in the mean <span class="hlt">field</span> signal with time resolution of 1 minute. Longer time scales readily show variations associated with rotation of magnetic patterns across the solar disk. Annual changes due to the varying visibility of the polar magnetic <span class="hlt">fields</span> may also be seen. Systematic effects associated with modulator non-uniformity require correction and limit the absolute accuracy of the GONG measurements. Comparison of the measurements with those from other instruments shows high correlation but suggest that GONG measurements of <span class="hlt">field</span> strength are low by a factor of about two. The source of this discrepancy is not clear. Fourier analysis of 2007 and 2008 time series of the GONG mean <span class="hlt">field</span> measurements shows strong signals at 27.75 and 26.84/2 day (synodic) periods with the later period showing more power. The heliospheric magnetic <span class="hlt">field</span> near Earth shows the same periods but with reversed power dominance. The Global Oscillation Network Group (GONG) project is managed by NSO, which is operated by AURA, Inc. under a cooperative agreement with the National Science Foundation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFMSM34B..04D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFMSM34B..04D"><span>Cluster <span class="hlt">observations</span> of Shear-mode surface waves diverging from Geomagnetic Tail reconnection</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dai, L.; Wygant, J. R.; Dombeck, J. P.; Cattell, C. A.; Thaller, S. A.; Mouikis, C.; Balogh, A.; Reme, H.</p> <p>2010-12-01</p> <p>We present the first Cluster spacecraft study of the intense (δB/B~0.5, δE/VAB~0.5) equatorial plane surface waves diverging from magnetic reconnection in the geomagnetic tail at ~17 Re. Using phase lag analysis with multi-spacecraft measurements, we quantitatively determine the wavelength and phase <span class="hlt">velocity</span> of the waves with spacecraft frame frequencies from 0.03 Hz to 1 Hz and wavelengths from much larger (4Re) than to comparable to the H+ gyroradius (~300km). The phase <span class="hlt">velocities</span> track the strong variations in the equatorial plane projection of the reconnection outflow <span class="hlt">velocity</span> perpendicular to the magnetic <span class="hlt">field</span>. The propagation direction and wavelength of the <span class="hlt">observed</span> surface waves resemble those of flapping waves of the magnetotail current sheet, suggesting a same origin shared by both of these waves. The <span class="hlt">observed</span> waves appear ubiquitous in the outflows near magnetotail reconnection. Evidence is found that the <span class="hlt">observed</span> waves are associated with <span class="hlt">velocity</span> shear in reconnection outflows. Analysis shows that <span class="hlt">observed</span> waves are associated with strong <span class="hlt">field</span>-aligned Alfvenic Poynting flux directed away from the reconnection region toward Earth. These <span class="hlt">observations</span> present a scenario in which the <span class="hlt">observed</span> surface waves are driven and convected through a <span class="hlt">velocity</span>-shear type instability by high-speed (~1000km) reconnection outflows tending to slow down due to power dissipation through Poynting flux. The mapped Poynting flux (100ergs/cm2s) and longitudinal scales (10-100 km) to 100km altitude suggest that the <span class="hlt">observed</span> waves and their motions are an important boundary condition for night-side aurora. Figure: a) The BX-GSM in the geomagnetic tail current sheet. b) The phase difference wavelet spectrum between Bz_GSM from SC2 and SC3, used to determine the wave phase <span class="hlt">velocity</span>, is correlated with the reconnection outflow <span class="hlt">velocity</span> (represented by H+ VX-GSM) c) The spacecraft trajectory through magnetotail reconnection. d) The <span class="hlt">observed</span> equatorial plane surface wave</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeoJI.213..931H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeoJI.213..931H"><span>Post-seismic <span class="hlt">velocity</span> changes following the 2010 Mw 7.1 Darfield earthquake, New Zealand, revealed by ambient seismic <span class="hlt">field</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>Heckels, R. EG; Savage, M. K.; Townend, J.</p> <p>2018-05-01</p> <p>Quantifying seismic <span class="hlt">velocity</span> changes following large earthquakes can provide insights into fault healing and reloading processes. This study presents temporal <span class="hlt">velocity</span> changes detected following the 2010 September Mw 7.1 Darfield event in Canterbury, New Zealand. We use continuous waveform data from several temporary seismic networks lying on and surrounding the Greendale Fault, with a maximum interstation distance of 156 km. Nine-component, day-long Green's functions were computed for frequencies between 0.1 and 1.0 Hz for continuous seismic records from immediately after the 2010 September 04 earthquake until 2011 January 10. Using the moving-window cross-spectral method, seismic <span class="hlt">velocity</span> changes were calculated. Over the study period, an increase in seismic <span class="hlt">velocity</span> of 0.14 ± 0.04 per cent was determined near the Greendale Fault, providing a new constraint on post-seismic relaxation rates in the region. A depth analysis further showed that <span class="hlt">velocity</span> changes were confined to the uppermost 5 km of the subsurface. We attribute the <span class="hlt">observed</span> changes to post-seismic relaxation via crack healing of the Greendale Fault and throughout the surrounding region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=teacher+AND+observation&pg=3&id=EJ1071938','ERIC'); return false;" href="https://eric.ed.gov/?q=teacher+AND+observation&pg=3&id=EJ1071938"><span>Rethinking <span class="hlt">Field</span> <span class="hlt">Observations</span>: Strengthening Teacher Education through INFORM</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>Hoyt, Kristin; Terantino, Joe</p> <p>2015-01-01</p> <p>This article introduces the Instructional <span class="hlt">Field</span> <span class="hlt">Observation</span> Rounds Model (INFORM), drawn from the medical profession where resident interns make rounds with experienced physicians, as an alternative approach for conducting classroom <span class="hlt">observations</span> in pre-service teacher education methods courses. INFORM centers on structured group <span class="hlt">observations</span> in…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeoJI.212..244D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeoJI.212..244D"><span>Resolving the fine-scale <span class="hlt">velocity</span> structure of continental hyperextension at the Deep Galicia Margin using full-waveform inversion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Davy, R. G.; Morgan, J. V.; Minshull, T. A.; Bayrakci, G.; Bull, J. M.; Klaeschen, D.; Reston, T. J.; Sawyer, D. S.; Lymer, G.; Cresswell, D.</p> <p>2018-01-01</p> <p>Continental hyperextension during magma-poor rifting at the Deep Galicia Margin is characterized by a complex pattern of faulting, thin continental fault blocks and the serpentinization, with local exhumation, of mantle peridotites along the S-reflector, interpreted as a detachment surface. In order to understand fully the evolution of these features, it is important to image seismically the structure and to model the <span class="hlt">velocity</span> structure to the greatest resolution possible. Traveltime tomography models have revealed the long-wavelength <span class="hlt">velocity</span> structure of this hyperextended domain, but are often insufficient to match accurately the short-wavelength structure <span class="hlt">observed</span> in reflection seismic imaging. Here, we demonstrate the application of 2-D time-domain acoustic full-waveform inversion (FWI) to deep-water seismic data collected at the Deep Galicia Margin, in order to attain a high-resolution <span class="hlt">velocity</span> model of continental hyperextension. We have used several quality assurance procedures to assess the <span class="hlt">velocity</span> model, including comparison of the <span class="hlt">observed</span> and modeled waveforms, checkerboard tests, testing of parameter and inversion strategy and comparison with the migrated reflection image. Our final model exhibits an increase in the resolution of subsurface <span class="hlt">velocities</span>, with particular improvement <span class="hlt">observed</span> in the westernmost continental fault blocks, with a clear rotation of the <span class="hlt">velocity</span> <span class="hlt">field</span> to match steeply dipping reflectors. Across the S-reflector, there is a sharpening in the <span class="hlt">velocity</span> contrast, with lower <span class="hlt">velocities</span> beneath S indicative of preferential mantle serpentinization. This study supports the hypothesis that normal faulting acts to hydrate the upper-mantle peridotite, <span class="hlt">observed</span> as a systematic decrease in seismic <span class="hlt">velocities</span>, consistent with increased serpentinization. Our results confirm the feasibility of applying the FWI method to sparse, deep-water crustal data sets.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19880009710','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19880009710"><span>Space vehicle approach <span class="hlt">velocity</span> judgments under simulated visual space conditions</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Haines, Richard F.</p> <p>1987-01-01</p> <p>Thirty-five volunteers responded when they first perceived an increase in apparent size of a collimated, 2-D image of an Orbiter vehicle. The test variables of interest included the presence of a fixed angular reticle within the <span class="hlt">field</span> of view (FOV); three initial Orbiter distances; three constant Orbiter approach <span class="hlt">velocities</span> corresponding to 1.6, 0.8, and 0.4 percent of the initial distance per second; and two background starfield <span class="hlt">velocities</span>. It was found that: (1) at each initial range, increasing approach <span class="hlt">velocity</span> led to a larger distance between the eye and Orbiter image at threshold; (2) including the fixed reticle in the FOV produced a smaller distance between the eye and Orbiter image at threshold; and (3) increasing background star <span class="hlt">velocity</span> during this judgment led to a smaller distance between the eye and Orbiter image at threshold. The last two findings suggest that other detail within the FOV may compete for available attention which otherwise would be available for judging image expansion; thus, the target has to approach the <span class="hlt">observer</span> nearer than otherwise if these details were present. These findings are discussed in relation to previous research and possible underlying mechanisms.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19770060049&hterms=method+magnetic&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dmethod%2Bmagnetic','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19770060049&hterms=method+magnetic&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dmethod%2Bmagnetic"><span>The mean magnetic <span class="hlt">field</span> of the sun - Method of <span class="hlt">observation</span> and relation to the interplanetary magnetic <span class="hlt">field</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Scherrer, P. H.; Wilcox, J. M.; Kotov, V.; Severnyi, A. B.; Howard, R.</p> <p>1977-01-01</p> <p>The mean solar magnetic <span class="hlt">field</span> as measured in integrated light has been <span class="hlt">observed</span> since 1968. Since 1970 it has been <span class="hlt">observed</span> both at Hale Observatories and at the Crimean Astrophysical Observatory. The <span class="hlt">observing</span> procedures at both observatories and their implications for mean <span class="hlt">field</span> measurements are discussed. A comparison of the two sets of daily <span class="hlt">observations</span> shows that similar results are obtained at both observatories. A comparison of the mean <span class="hlt">field</span> with the interplanetary magnetic polarity shows that the IMF sector structure has the same pattern as the mean <span class="hlt">field</span> polarity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015HiA....16..399K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015HiA....16..399K"><span>Magnetic <span class="hlt">fields</span> in spiral galaxies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Krause, Marita</p> <p>2015-03-01</p> <p>The magnetic <span class="hlt">field</span> structure in edge-on galaxies <span class="hlt">observed</span> so far shows a plane-parallel magnetic <span class="hlt">field</span> component in the disk of the galaxy and an X-shaped <span class="hlt">field</span> in its halo. The plane-parallel <span class="hlt">field</span> is thought to be the projected axisymmetric (ASS) disk <span class="hlt">field</span> as <span class="hlt">observed</span> in face-on galaxies. Some galaxies addionionally exhibit strong vertical magnetic <span class="hlt">fields</span> in the halo right above and below the central region of the disk. The mean-<span class="hlt">field</span> dynamo theory in the disk cannot explain these <span class="hlt">observed</span> <span class="hlt">fields</span> without the action of a wind, which also probably plays an important role to keep the vertical scale heights constant in galaxies of different Hubble types and star formation activities, as has been <span class="hlt">observed</span> in the radio continuum: At λ6 cm the vertical scale heights of the thin disk and the thick disk/halo in a sample of five edge-on galaxies are similar with a mean value of 300 +/- 50 pc for the thin disk and 1.8 +/- 0.2 kpc for the thick disk (a table and references are given in Krause 2011) with our sample including the brightest halo <span class="hlt">observed</span> so far, NGC 253, with strong star formation, as well as one of the weakest halos, NGC 4565, with weak star formation. If synchrotron emission is the dominant loss process of the relativistic electrons the outer shape of the radio emission should be dumbbell-like as has been <span class="hlt">observed</span> in several edge-on galaxies like e.g. NGC 253 (Heesen et al. 2009) and NGC 4565. As the synchrotron lifetime t syn at a single frequency is proportional to the total magnetic <span class="hlt">field</span> strength B t -1.5, a cosmic ray bulk speed (<span class="hlt">velocity</span> of a galactic wind) can be defined as v CR = h CR /t syn = 2 h z /t syn , where h CR and h z are the scale heights of the cosmic rays and the <span class="hlt">observed</span> radio emission at this freqnency. Similar <span class="hlt">observed</span> radio scale heights imply a self regulation mechanism between the galactic wind <span class="hlt">velocity</span>, the total magnetic <span class="hlt">field</span> strength and the star formation rate SFR in the disk: v CR ~ B t 1.5 ~ SFR ~ 0.5 (Niklas & Beck 1997).</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>