To see the other types of publications on this topic, follow the link: Transverse velocity.

Journal articles on the topic 'Transverse velocity'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Transverse velocity.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Jensen, Jørgen. "Transverse spectral velocity estimation." IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control 61, no. 11 (November 2014): 1815–23. http://dx.doi.org/10.1109/tuffc.2014.006488.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Andreu-Angulo, Ignacio, Holger Babinsky, Hülya Biler, Girguis Sedky, and Anya R. Jones. "Effect of Transverse Gust Velocity Profiles." AIAA Journal 58, no. 12 (December 2020): 5123–33. http://dx.doi.org/10.2514/1.j059665.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Wang, Hao, Jie Mao, Ke Liu, Sheng Wang, and Liang Yu. "Transverse Velocity Effect on Hunt’s Flow." IEEE Transactions on Plasma Science 46, no. 5 (May 2018): 1534–38. http://dx.doi.org/10.1109/tps.2017.2777844.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

De Colle, Fabio, Adriano H. Cerqueira, and Angels Riera. "TRANSVERSE VELOCITY SHIFTS IN PROTOSTELLAR JETS: ROTATION OR VELOCITY ASYMMETRIES?" Astrophysical Journal 832, no. 2 (November 29, 2016): 152. http://dx.doi.org/10.3847/0004-637x/832/2/152.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Lin, Hong, Chang Li Zhou, Jun Shi, and Zhi Hua Feng. "Transverse Vibration of Axially Accelerating Moving Fabric: Experiment and Analysis." Applied Mechanics and Materials 226-228 (November 2012): 150–53. http://dx.doi.org/10.4028/www.scientific.net/amm.226-228.150.

Full text
Abstract:
Axially moving fabric can be met in many textile devices. In most cases, the transverse vibrations of fabric can cause a series of negative influents to the product. In this paper, the transverse vibration of axially accelerating moving fabric, which is excited by velocity fluctuations, is investigated by experimental method. The harmonic varying velocity is achieved through a brushless DC motor controlled by PWM technology based on the embedded microcontroller LPC1768. An inductive non-contact displacement sensor is used to measure transversal vibration of fabric. The motor speed is measured by a photoelectric encoder. The experimental data is processed by measurement platform based on Labview and the analysis is given. Laboratory measurements demonstrate the effect of velocity fluctuations on transverse vibration of fabric, particularly near the parametric resonance region.
APA, Harvard, Vancouver, ISO, and other styles
6

Itoh, Naoki, and Takemi Kotouda. "Velocity-Magnetic Field Correlation of Pulsars." International Astronomical Union Colloquium 160 (1996): 49–50. http://dx.doi.org/10.1017/s0252921100040999.

Full text
Abstract:
Monte Carlo simulations of the evolution of pulsars are carried out in order to compare with the recent measurement of the pulsar transverse velocity by Lyne & Lorimer (1994). The new electron density distribution model of Taylor & Cordes (1993) is adopted in the simulation. Accurate pulsar orbits in the Galactic gravitational field are calculated. It is found that the constant magnetic field model of pulsars can account for the new measurement of the pulsar transverse velocity and the apparent correlation between the strength of the magnetic field and the transverse velocity of the pulsars. The present finding confirms the validity of the constant magnetic field model of pulsars and consolidates the idea that the apparent correlation between the strength of the magnetic field and the transverse velocity of the pulsars is caused by observational selection effects.
APA, Harvard, Vancouver, ISO, and other styles
7

Ristić, D., Mile Ivanda, K. Furić, M. Montagna, Maurizio Ferrari, A. Chiasera, and Yoann Jestin. "Raman Scattering on the l=2 Spheroidal Mode of Spherical Nanoparticles." Advances in Science and Technology 55 (September 2008): 132–37. http://dx.doi.org/10.4028/www.scientific.net/ast.55.132.

Full text
Abstract:
The Raman light to vibrations coupling coefficients C(ν) of quadrupolar and symmetrical vibrational modes of spherical nanoparticles embedded in a matrix are calculated. In contrast to the symmetrical mode, the C(ν) of the quadrupolar modes consists of the longitudinal and transversal sound velocity contributions. It is shown, that depending on the ratio of longitudinal and transverse sound velocity, these two contributions can interfere constructively or destructively resulting in enhancing or vanishing of some radial modes. Different peaks in the C(ν) spectrum were attributed to transverse and longitudinal spheroidal modes and the longitudinal spheroidal modes were found to have a higher Raman intensity than the transverse modes. The theoretical model was tested on a sample of HfO2 nanoparticles in a silica matrix.
APA, Harvard, Vancouver, ISO, and other styles
8

Cohen, Jack K. "Analytic study of the effective parameters for determination of the NMO velocity function in transversely isotropic media." GEOPHYSICS 62, no. 6 (November 1997): 1855–66. http://dx.doi.org/10.1190/1.1444286.

Full text
Abstract:
In their studies of transversely isotropic media with a vertical symmetry axis (VTI media), Alkhalifah and Tsvankin observed that, to a high numerical accuracy, the normal moveout (NMO) velocity for dipping reflectors as a function of ray parameter p depends mainly on just two parameters, each of which can be determined from surface P‐wave observations. They substantiated this result by using the weak‐anisotropy approximation and exploited it to develop a time‐domain processing sequence that takes into account vertical transverse isotropy. In this study, the two‐parameter Alkhalifah‐Tsvankin result was further examined analytically. It was found that although there is (as these authors already observed) some dependence on the remaining parameters of the problem, this dependence is weak, especially in the practically important regimes of weak to moderately strong transverse isotropy and small ray parameter. In each of these regimes, an analytic solution is derived for the anisotropy parameter η required for time‐domain P‐wave imaging in VTI media. In the case of elliptical anisotropy (η = 0), NMO velocity expressed through p is fully controlled just by the zero‐dip NMO velocity—one of the Alkhalifah‐ Tsvankin parameters. The two‐parameter representation of NMO velocity also was shown to be exact in another limit—that of the zero shear‐wave vertical velociy. The analytic results derived here are based on new representations for both the P‐wave phase velocity and normal moveout velocity in terms of the ray parameter, with explicit expressions given for the cases of vanishing onaxis shear speed, weak to moderate transverse isotropy, and small to moderate ray parameter. Using these formulas, I have rederived and, in some cases, extended in a uniform manner various results of Tsvankin, Alkhalifah, and others. Examples include second‐order expansions in the anisotropy parameters for both the P‐wave phase‐velocity function and NMO‐velocity function, as well as expansions in powers of the ray parameter for both of these functions. I have checked these expansions against the corresponding exact functions for several choices of the anisotropy parameters.
APA, Harvard, Vancouver, ISO, and other styles
9

Yan, Ting-Zhi, and Shan Li. "Transverse radius dependence for transverse velocity and elliptic flow in intermediate energy HIC." Chinese Physics C 35, no. 5 (May 2011): 459–62. http://dx.doi.org/10.1088/1674-1137/35/5/010.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Amarante, João A. S., Martin C. Smith, and Corrado Boeche. "The high transverse velocity stars in Gaia-LAMOST." Proceedings of the International Astronomical Union 14, S353 (June 2019): 59–60. http://dx.doi.org/10.1017/s1743921319008603.

Full text
Abstract:
AbstractAlthough the stellar halo accounts for just ∼1% of the total stellar mass of the Milky Way, the kinematics of halo stars can tell us a lot about the origins and evolution of our Galaxy. It has been shown that the high transverse velocity stars in Gaia DR2 reveal a double sequence in the Hertzsprung-Russell (HR) diagram, indicating a duality in the local halo within 1 kpc. We fit these stars by updating the popular Besançon/Galaxia model, incorporating the latest observational results for the stellar halo. We are able to obtain a good match to the Gaia data and provide new constraints on the properties of the disc and halo. In particular, we show that the thick disc contribution to this high velocity tail is small, but not negligible, and likely has an influence on the red sequence of the HR diagram.
APA, Harvard, Vancouver, ISO, and other styles
11

Xu, G., R. A. Antonia, and S. Rajagopalan. "Scaling of mixed longitudinal-transverse velocity structure functions." Europhysics Letters (EPL) 79, no. 4 (July 27, 2007): 44001. http://dx.doi.org/10.1209/0295-5075/79/44001.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

De Moortel, I., D. J. Pascoe, A. N. Wright, and A. W. Hood. "Transverse, propagating velocity perturbations in solar coronal loops." Plasma Physics and Controlled Fusion 58, no. 1 (October 20, 2015): 014001. http://dx.doi.org/10.1088/0741-3335/58/1/014001.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Leng, Xinqian, and Hubert Chanson. "Transverse velocity profiling under positive surges in channels." Flow Measurement and Instrumentation 64 (December 2018): 14–27. http://dx.doi.org/10.1016/j.flowmeasinst.2018.10.006.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Nizamova, A. D., V. N. Kireev, and S. F. Urmancheev. "Research of eigenfuctions perturbation of the transverse component velocity thermoviscous liquids flow." Multiphase Systems 14, no. 2 (2019): 132–37. http://dx.doi.org/10.21662/mfs2019.2.018.

Full text
Abstract:
The viscous model fluid flow in a plane channel with a linear temperature profile is considered. The problem of the thermoviscous fluid flow stability is solved on the basis of the previously obtained generalized Orr–Sommerfeld equation by the spectral method of decomposition into Chebyshev polynomials. We study the effect of taking into account the linear and exponential dependences of the viscosity of a liquid on temperature on the eigenfunctions of the hydrodynamic stability equation and on perturbations of the transverse velocity of an incompressible fluid in a plane channel when various wall temperatures are specified. Eigenfunctions are found numerically for two eigenvalues of the linear and exponential dependence of viscosity on temperature. Presented pictures of their own functions. The eigenfunctions demonstrate the behavior of the transverse velocity perturbations, their possible growth or attenuation over time. For the given eigenfunctions, perturbations of the transverse flow velocity of a thermoviscous fluid are obtained. It is shown that taking the temperature dependence of viscosity into account affects the eigenfunctions of the equations of hydrodynamic stability and perturbations of the transverse flow velocity. Perturbations of the transverse velocity significantly affect the hydrodynamic instability of the fluid flow. The results show that when considering the unstable eigenvalue over time, the velocity perturbations begin to grow, which leads to turbulence of the flow. The maximum values of the eigenfunctions and perturbations of the transverse velocities are shifted to the hot wall. It is seen that for an unstable eigenvalue, the perturbations of the transverse flow velocity increase over time, and for a stable one, they decay.
APA, Harvard, Vancouver, ISO, and other styles
15

Tsvankin, Ilya, and Leon Thomsen. "Inversion of reflection traveltimes for transverse isotropy." GEOPHYSICS 60, no. 4 (July 1995): 1095–107. http://dx.doi.org/10.1190/1.1443838.

Full text
Abstract:
In anisotropic media, the short‐spread stacking velocity is generally different from the root‐mean‐square vertical velocity. The influence of anisotropy makes it impossible to recover the vertical velocity (or the reflector depth) using hyperbolic moveout analysis on short‐spread, common‐midpoint (CMP) gathers, even if both P‐ and S‐waves are recorded. Hence, we examine the feasibility of inverting long‐spread (nonhyperbolic) reflection moveouts for parameters of transversely isotropic media with a vertical symmetry axis. One possible solution is to recover the quartic term of the Taylor series expansion for [Formula: see text] curves for P‐ and SV‐waves, and to use it to determine the anisotropy. However, this procedure turns out to be unstable because of the ambiguity in the joint inversion of intermediate‐spread (i.e., spreads of about 1.5 times the reflector depth) P and SV moveouts. The nonuniqueness cannot be overcome by using long spreads (twice as large as the reflector depth) if only P‐wave data are included. A general analysis of the P‐wave inverse problem proves the existence of a broad set of models with different vertical velocities, all of which provide a satisfactory fit to the exact traveltimes. This strong ambiguity is explained by a trade‐off between vertical velocity and the parameters of anisotropy on gathers with a limited angle coverage. The accuracy of the inversion procedure may be significantly increased by combining both long‐spread P and SV moveouts. The high sensitivity of the long‐spread SV moveout to the reflector depth permits a less ambiguous inversion. In some cases, the SV moveout alone may be used to recover the vertical S‐wave velocity, and hence the depth. Success of this inversion depends on the spreadlength and degree of SV‐wave velocity anisotropy, as well as on the constraints on the P‐wave vertical velocity.
APA, Harvard, Vancouver, ISO, and other styles
16

Dong, Zhi Yong, and Guo Liang Yu. "A Preliminary Study of Effects of Wave and Crossflow on Side Discharge of Wastewater." Advanced Materials Research 955-959 (June 2014): 2041–45. http://dx.doi.org/10.4028/www.scientific.net/amr.955-959.2041.

Full text
Abstract:
Transverse and longitudinal velocity profiles on the xoy plane at z=0 and 5cm of side discharge jet under the action of crossflow and wave, velocity vectors on the xoy plane at z=0 and 5cm of side discharge jet under the action of wave and crossflow were measured by Micro ADV in this paper, respectively. Effects of wave and crossflow on velocity profiles and velocity vectors of side discharge jet were analyzed. The preliminary experimental results showed that transverse velocity profiles of side discharge jet centerline was less affected by crossflow and wave, however, effects on transverse and longitudinal velocities over the jet centerline were significant. Also, velocity vectors over of side discharge jet centerline were remarkably affected by wave and crossflow though velocity vectors on the jet centerline was less affected.
APA, Harvard, Vancouver, ISO, and other styles
17

Tsvankin, Ilya. "P‐wave signatures and notation for transversely isotropic media: An overview." GEOPHYSICS 61, no. 2 (March 1996): 467–83. http://dx.doi.org/10.1190/1.1443974.

Full text
Abstract:
Progress in seismic inversion and processing in anisotropic media depends on our ability to relate different seismic signatures to the anisotropic parameters. While the conventional notation (stiffness coefficients) is suitable for forward modeling, it is inconvenient in developing analytic insight into the influence of anisotropy on wave propagation. Here, a consistent description of P‐wave signatures in transversely isotropic (TI) media with arbitrary strength of the anisotropy is given in terms of Thomsen notation. The influence of transverse isotropy on P‐wave propagation is shown to be practically independent of the vertical S‐wave velocity [Formula: see text], even in models with strong velocity variations. Therefore, the contribution of transverse isotropy to P‐wave kinematic and dynamic signatures is controlled by just two anisotropic parameters, ε and δ, with the vertical velocity [Formula: see text] being a scaling coefficient in homogeneous models. The distortions of reflection moveouts and amplitudes are not necessarily correlated with the magnitude of velocity anisotropy. The influence of transverse isotropy on P‐wave normal‐moveout (NMO) velocity in a horizontally layered medium, on small‐angle reflection coefficient, and on point‐force radiation in the symmetry direction is entirely determined by the parameter δ. Another group of signatures of interest in reflection seisimology—the dip‐dependence of NMO velocity, magnitude of nonhyperbolic moveout, time‐migration impulse response, and the radiation pattern near vertical—is dependent on both anisotropic parameters (ε and δ) and is primarily governed by the difference between ε and δ. Since P‐wave signatures are so sensitive to the value of ε − δ, application of the elliptical‐anisotropy approximation (ε = δ) in P‐wave processing may lead to significant errors. Many analytic expressions given in the paper remain valid in transversely isotropic media with a tilted symmetry axis. Moreover, the equation for NMO velocity from dipping reflectors, as well as the nonhyperbolic moveout equation, can be used in symmetry planes of any anisotropic media (e.g., orthorhombic).
APA, Harvard, Vancouver, ISO, and other styles
18

Sundarnath, J. K., and R. Muthucumarswamy. "Hall Effects on Mhd Flow Past an Accelerated Plate with Heat Transfer." International Journal of Applied Mechanics and Engineering 20, no. 1 (February 1, 2015): 171–81. http://dx.doi.org/10.1515/ijame-2015-0012.

Full text
Abstract:
Abstract Hall current and rotation on an MHD flow past an accelerated horizontal plate relative to a rotating fluid, in the presence of heat transfer has been analyzed. The effects of the Hall parameter, Hartmann number, rotation parameter (non-dimensional angular velocity), Grashof’s number and Prandtl number on axial and transverse velocity profiles are presented graphically. It is found that with the increase in the Hartmann number, the axial and transverse velocity components increase in a direction opposite to that of obtained by increasing the Hall parameter and rotation parameter. Also, when Ω=M2m /(1 + m2 ) , it is observed that the transverse velocity component vanishes and axial velocity attains a maximum value.
APA, Harvard, Vancouver, ISO, and other styles
19

He, Yue-Lei, Jian-Kang Shen, Zai-Wei Li, and Hong-Yao Lu. "Fractal Characteristics of Transverse Crack Propagation on CRTSII Type Track Slab." Mathematical Problems in Engineering 2019 (July 25, 2019): 1–9. http://dx.doi.org/10.1155/2019/6587343.

Full text
Abstract:
Objectives of this study are to examine influence of the surface crackles fractal characteristics of CRTSII track slab on its propagation velocity and stress intensity factor (SIF) and to improve the service performance of track slab. In this paper, fractal dimension D of transverse crack on the surface of CRTSII track slab was calculated by image processing, digital image technology, and box-counting method. Based on the fractal structure of transverse crack on track slab surface, a fractal transverse crack propagation model was established. The impact of fractal effect on the surface crack propagation velocity and SIF of CRTSII track slab was analyzed quantitatively. The fractal dimension D of transverse crack is 1.0652 on the surface of CRTSII track slab. The actual measured crack propagation velocity V0 on the surface of the track slab can be corrected as the true crack propagation velocity V. When the V0/Cr=0.899<1, the SIF K′L1.0652,t,V of transverse dynamic fractal crack propagation on the surface of track slab has been equal to 0, and the V0 is always lower than Rayleigh wave velocity. In addition, the true attenuation rate of SIF for CRTSII track slab surface transverse dynamic crack propagation is faster than the actual measured value.
APA, Harvard, Vancouver, ISO, and other styles
20

Han, Wenhu, Cheng Wang, and Chung K. Law. "Role of transversal concentration gradient in detonation propagation." Journal of Fluid Mechanics 865 (February 22, 2019): 602–49. http://dx.doi.org/10.1017/jfm.2019.37.

Full text
Abstract:
The role of a transversal concentration gradient in detonation propagation in a two-dimensional channel filled with an $\text{H}_{2}{-}\text{O}_{2}$ mixture is examined by high-resolution simulation. Results show that, compared to propagation in homogeneous media, a concentration gradient reduces the average detonation velocity because of the delay in reaching downstream reaction equilibrium, leading to a large amount of unreacted $\text{H}_{2}$ and hence significant species fluctuations. The transversal concentration gradient also enhances the cellular detonation instability. Steepening it reduces considerably the number of triple points on the front, lengthens the global detonation front structure on average and consequently increases the deficit of the average detonation velocity. It is further found that the interaction of the leading shock with the transversal concentration gradient influences the formation of local $\text{H}_{2}$ bump and thus the unreacted pocket behind the front, while the transverse wave causes mixing and burning of the residue fuel downstream. Nevertheless, for the steepened concentration gradient, a transverse detonation is present and consumes the fuel in the compressed and preheated zone by the leading shock; consequently, the detonation velocity deficit is not increased significantly for detonation with the single-head propagation mode close to the limit.
APA, Harvard, Vancouver, ISO, and other styles
21

PEARSON, B. R., and R. A. ANTONIA. "Reynolds-number dependence of turbulent velocity and pressure increments." Journal of Fluid Mechanics 444 (September 25, 2001): 343–82. http://dx.doi.org/10.1017/s0022112001005511.

Full text
Abstract:
The main focus is the Reynolds number dependence of Kolmogorov normalized low-order moments of longitudinal and transverse velocity increments. The velocity increments are obtained in a large number of flows and over a wide range (40–4250) of the Taylor microscale Reynolds number Rλ. The Rλ dependence is examined for values of the separation, r, in the dissipative range, inertial range and in excess of the integral length scale. In each range, the Kolmogorov-normalized moments of longitudinal and transverse velocity increments increase with Rλ. The scaling exponents of both longitudinal and transverse velocity increments increase with Rλ, the increase being more significant for the latter than the former. As Rλ increases, the inequality between scaling exponents of longitudinal and transverse velocity increments diminishes, reflecting a reduced influence from the large-scale anisotropy or the mean shear on inertial range scales. At sufficiently large Rλ, inertial range exponents for the second-order moment of the pressure increment follow more closely those for the fourth-order moments of transverse velocity increments than the fourth-order moments of longitudinal velocity increments. Comparison with DNS data indicates that the magnitude and Rλ dependence of the mean square pressure gradient, based on the joint-Gaussian approximation, is incorrect. The validity of this approximation improves as r increases; when r exceeds the integral length scale, the Rλ dependence of the second-order pressure structure functions is in reasonable agreement with the result originally given by Batchelor (1951).
APA, Harvard, Vancouver, ISO, and other styles
22

Fard, Mehrdad P., and Svein I. Sagatun. "Exponential Stabilization of a Transversely Vibrating Beam by Boundary Control Via Lyapunov’s Direct Method." Journal of Dynamic Systems, Measurement, and Control 123, no. 2 (November 2, 1999): 195–200. http://dx.doi.org/10.1115/1.1369111.

Full text
Abstract:
This paper discusses the boundary stabilization of a beam in free transverse vibration. The dynamics of the beam is presented by a nonlinear partial differential equation (PDE). Based on this model a nonlinear control law is constructed to stabilize the system. The control law is a nonlinear function of the slopes and velocity at the boundary of the beam. The novelty of this article is that it has been possible to exponentially stabilize a free transversely vibrating beam via boundary control without restoring to truncation of the model. This result is achieved while the coupling between longitudinal and transversal displacements has been taken into account.
APA, Harvard, Vancouver, ISO, and other styles
23

URBAN, WILLIAM D., and M. G. MUNGAL. "Planar velocity measurements in compressible mixing layers." Journal of Fluid Mechanics 431 (March 25, 2001): 189–222. http://dx.doi.org/10.1017/s0022112000003177.

Full text
Abstract:
In the present study, we use particle image velocimetry (PIV) to obtain planar, two-component velocity fields in two-dimensional, turbulent mixing layers at convective Mach numbers Mc of 0.25, 0.63, and 0.76. The experiments are performed in a large-scale blowdown wind tunnel, with high-speed free-stream Mach numbers up to 2.25 and shear-layer Reynolds numbers up to 106. Specific issues relating to the application of PIV to supersonic flows are addressed. The instantaneous data are analysed to produce maps of derived quantities such as vorticity, and ensemble averaged to provide turbulence statistics. The results show that compressibility introduces marked changes in the disposition of instantaneous velocity gradients within the layer, and hence in the vorticity field. In particular, peak transverse vorticity values are seen to be confined to thin streamwise sheets under compressible conditions, with little transverse communication. The location of these sheets near the lab-frame sonic velocity suggests a sensitivity of the compressible layer to stationary disturbances. Turbulence statistics derived from the planar velocity measurements confirm previous observations of strong suppression of transverse velocity fluctuations and primary Reynolds stress as Mc increases between 0.25 and 0.76.
APA, Harvard, Vancouver, ISO, and other styles
24

Al‐Dajani, AbdulFattah, and Ilya Tsvankin. "Nonhyperbolic reflection moveout for horizontal transverse isotropy." GEOPHYSICS 63, no. 5 (September 1998): 1738–53. http://dx.doi.org/10.1190/1.1444469.

Full text
Abstract:
The transversely isotropic model with a horizontal axis of symmetry (HTI) has been used extensively in studies of shear‐wave splitting to describe fractured formations with a single system of parallel vertical penny‐shaped cracks. Here, we present an analytic description of longspread reflection moveout in horizontally layered HTI media with arbitrary strength of anisotropy. The hyperbolic moveout equation parameterized by the exact normal‐moveout (NMO) velocity is sufficiently accurate for P-waves on conventional‐length spreads (close to the reflector depth), although the NMO velocity is not, in general, usable for converting time to depth. However, the influence of anisotropy leads to the deviation of the moveout curve from a hyperbola with increasing spread length, even in a single‐layer model. To account for nonhyperbolic moveout, we have derived an exact expression for the azimuthally dependent quartic term of the Taylor series traveltime expansion [t2(x2)] valid for any pure mode in an HTI layer. The quartic moveout coefficient and the NMO velocity are then substituted into the nonhyperbolic moveout equation of Tsvankin and Thomsen, originally designed for vertical transverse isotropy (VTI). Numerical examples for media with both moderate and uncommonly strong nonhyperbolic moveout show that this equation accurately describes azimuthally dependent P-wave reflection traveltimes in an HTI layer, even for spread lengths twice as large as the reflector depth. In multilayered HTI media, the NMO velocity and the quartic moveout coefficient reflect the influence of layering as well as azimuthal anisotropy. We show that the conventional Dix equation for NMO velocity remains entirely valid for any azimuth in HTI media if the group‐velocity vectors (rays) for data in a common‐midpoint (CMP) gather do not deviate from the vertical incidence plane. Although this condition is not exactly satisfied in the presence of azimuthal velocity variations, rms averaging of the interval NMO velocities represents a good approximation for models with moderate azimuthal anisotropy. Furthermore, the quartic moveout coefficient for multilayered HTI media can also be calculated with acceptable accuracy using the known averaging equations for vertical transverse isotropy. This allows us to extend the nonhyperbolic moveout equation to horizontally stratified media composed of any combination of isotropic, VTI, and HTI layers. In addition to providing analytic insight into the behavior of nonhyperbolic moveout, these results can be used in modeling and inversion of reflection traveltimes in azimuthally anisotropic media.
APA, Harvard, Vancouver, ISO, and other styles
25

Tsvankin, Ilya. "Normal moveout from dipping reflectors in anisotropic media." GEOPHYSICS 60, no. 1 (January 1995): 268–84. http://dx.doi.org/10.1190/1.1443755.

Full text
Abstract:
Description of reflection moveout from dipping interfaces is important in developing seismic processing methods for anisotropic media, as well as in the inversion of reflection data. Here, I present a concise analytic expression for normal‐moveout (NMO) velocities valid for a wide range of homogeneous anisotropic models including transverse isotropy with a tilted in‐plane symmetry axis and symmetry planes in orthorhombic media. In transversely isotropic media, NMO velocity for quasi‐P‐waves may deviate substantially from the isotropic cosine‐of‐dip dependence used in conventional constant‐velocity dip‐moveout (DMO) algorithms. However, numerical studies of NMO velocities have revealed no apparent correlation between the conventional measures of anisotropy and errors in the cosine‐of‐dip DMO correction (“DMO errors”). The analytic treatment developed here shows that for transverse isotropy with a vertical symmetry axis, the magnitude of DMO errors is dependent primarily on the difference between Thomsen parameters ε and δ. For the most common case, ε − δ > 0, the cosine‐of‐dip–corrected moveout velocity remains significantly larger than the moveout velocity for a horizontal reflector. DMO errors at a dip of 45 degrees may exceed 20–25 percent, even for weak anisotropy. By comparing analytically derived NMO velocities with moveout velocities calculated on finite spreads, I analyze anisotropy‐induced deviations from hyperbolic moveout for dipping reflectors. For transversely isotropic media with a vertical velocity gradient and typical (positive) values of the difference ε − δ, inhomogeneity tends to reduce (sometimes significantly) the influence of anisotropy on the dip dependence of moveout velocity.
APA, Harvard, Vancouver, ISO, and other styles
26

Waqas, Muhammad, and Guang-Xiong Peng. "Study of Dependence of Kinetic Freezeout Temperature on the Production Cross-Section of Particles in Various Centrality Intervals in Au–Au and Pb–Pb Collisions at High Energies." Entropy 23, no. 4 (April 20, 2021): 488. http://dx.doi.org/10.3390/e23040488.

Full text
Abstract:
Transverse momentum spectra of π+, p, Λ, Ξ or Ξ¯+, Ω or Ω¯+ and deuteron (d) in different centrality intervals in nucleus–nucleus collisions at the center of mass energy are analyzed by the blast wave model with Boltzmann Gibbs statistics. We extracted the kinetic freezeout temperature, transverse flow velocity and kinetic freezeout volume from the transverse momentum spectra of the particles. It is observed that the non-strange and strange (multi-strange) particles freezeout separately due to different reaction cross-sections. While the freezeout volume and transverse flow velocity are mass dependent, they decrease with the resting mass of the particles. The present work reveals the scenario of a double kinetic freezeout in nucleus–nucleus collisions. Furthermore, the kinetic freezeout temperature and freezeout volume are larger in central collisions than peripheral collisions. However, the transverse flow velocity remains almost unchanged from central to peripheral collisions.
APA, Harvard, Vancouver, ISO, and other styles
27

Beck, R. Anne, and James M. Longuski. "Annihilation of Transverse Velocity Bias During Spinning-Up Maneuvers." Journal of Guidance, Control, and Dynamics 20, no. 3 (May 1997): 416–21. http://dx.doi.org/10.2514/2.4074.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Wyithe, J. S. B., R. L. Webster, and E. L. Turner. "A measurement of the transverse velocity of Q2237+0305." Monthly Notices of the Royal Astronomical Society 309, no. 1 (October 11, 1999): 261–72. http://dx.doi.org/10.1046/j.1365-8711.1999.02844.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Li, Hui-Quan, Jian-Cheng Wang, and Li Xue. "The Synchrotron Emission of Jets with Transverse Velocity Discrepancy." Chinese Journal of Astronomy and Astrophysics 4, no. 4 (August 2004): 311–19. http://dx.doi.org/10.1088/1009-9271/4/4/311.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Yasini, Siavash, Nareg Mirzatuny, and Elena Pierpaoli. "Pairwise Transverse Velocity Measurement with the Rees–Sciama Effect." Astrophysical Journal 873, no. 2 (March 15, 2019): L23. http://dx.doi.org/10.3847/2041-8213/ab0bfe.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Hirschowitz, P. M., and C. S. James. "Transverse velocity distributions in channels with emergent bank vegetation." River Research and Applications 25, no. 9 (November 2009): 1177–92. http://dx.doi.org/10.1002/rra.1216.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Ganguli, G., Y. C. Lee, and P. J. Palmadesso. "Electron–ion hybrid mode due to transverse velocity shear." Physics of Fluids 31, no. 10 (1988): 2753. http://dx.doi.org/10.1063/1.866982.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Chen, Guanghua, Detian Wang, Jun Liu, Jianhua Meng, Shouxian Liu, and Qingguo Yang. "A novel photonic Doppler velocimetry for transverse velocity measurement." Review of Scientific Instruments 84, no. 1 (January 2013): 013101. http://dx.doi.org/10.1063/1.4776186.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Zhou, T., R. A. Antonia, J.-J. Lasserre, M. Coantic, and F. Anselmet. "Transverse velocity and temperature derivative measurements in grid turbulence." Experiments in Fluids 34, no. 4 (February 7, 2003): 449–59. http://dx.doi.org/10.1007/s00348-002-0566-9.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Yi, Cheng Wu, Shuai Ma, Yun Qing Zhao, and Rong Jie Yi. "Study on Ion Concentration of New Type Transverse Plate ESP." Advanced Materials Research 468-471 (February 2012): 2381–84. http://dx.doi.org/10.4028/www.scientific.net/amr.468-471.2381.

Full text
Abstract:
According to the transverse plate electrostatic precipitator with high velocity collect dust theory, established the laboratory scale transverse plate ESP combines the hydrodynamic, static electrics. In this paper, experiment of ion concentration are carried using the transverse plate ESP. (The laboratory scale transverse plate electrostatic precipitator self-designed) system. The influence rules of the factors to ion concentration are examined such as distance between the export, the discharge electrode, applied voltage, internal of dust collection plates and the gas velocity. According to the experiment result, the ion concentration can increase about 109/cm3, when the working voltage is 18kV, the gas velocity is 4m/s, the distance of effective dust collecting plate is 40mm. The ion concentration of electrostatics precipitator system reaches the maximum and is above 109.
APA, Harvard, Vancouver, ISO, and other styles
36

Liu, Qiang, and Lin-jing Xiao. "Comparative Analysis of Longitudinal and Transverse Vibration Characteristics of Ocean Mining Pipe." Shock and Vibration 2021 (June 11, 2021): 1–25. http://dx.doi.org/10.1155/2021/5546371.

Full text
Abstract:
In this paper, the 5000 m mining pipe is taken as the research object, and the transverse and longitudinal vibration laws of the pipe under different working conditions are analyzed. Based on the finite element method (FEM), the pipe is discretized and calculated by the Wilson-θ Wilson - θ integral method; finally, the corresponding vibration laws of the mining pipe are obtained. The research shows that the mining pipe vibration responses are irregular motion, with the obvious oscillation phenomenon, and the overall vibration trend decreases first and then increases from the top to the bottom; the maximum vibration response occurs at the pipe top. Under the same working conditions, increasing the towing velocity will decrease the overall longitudinal vibration amplitude and increase the overall transverse vibration amplitude. While the ore bin weight will increase the longitudinal vibration amplitude and decrease the transverse vibration amplitude, increasing the mining pipe large diameter stepped section length and damping will decrease the longitudinal and transverse vibration simultaneously. When the towing velocity is between 0–2.8 m/s, the longitudinal vibration intensity is large, which is the main vibration mode. When the towing velocity is 2.8 m/s, the critical point is reached, and the longitudinal and transverse vibrations have the same intensity. When the towing velocity is greater than 2.8 m/s, the transverse vibration intensity is gradually greater than the longitudinal vibration intensity; at this time, the control of the transverse vibration should be appropriately increased.
APA, Harvard, Vancouver, ISO, and other styles
37

ZHANG, XIN, and FUGEN WU. "STRONG MODE TRANSFORMATION OF ELASTIC WAVES BY SPHERICAL SOLID OBJECT." Modern Physics Letters B 23, no. 17 (July 10, 2009): 2155–65. http://dx.doi.org/10.1142/s0217984909020308.

Full text
Abstract:
We investigate mode transformation between longitudinal mode and transverse mode by spherical solid object. By analyzing the scattering cross section, we find that longitudinal wave (L) transformation to transverse wave (T) is much more than transverse wave (T) transformation to longitudinal wave (L). The ratio of L → T converted waves to T → L converted waves is verified to be independent of frequency. The strongest mode transformation between transverse mode and longitudinal mode can be obtained by choosing a solid sphere with large density and low shear velocity and background with small density and high shear velocity.
APA, Harvard, Vancouver, ISO, and other styles
38

Mykhailo, Piven. "RESEARCH ON THE EFFECT OF THE INITIAL SPEED OF THE MIXTURE ON THE PROCESS OF LOADING." Vibrations in engineering and technology, no. 4(95) (December 20, 2019): 47–55. http://dx.doi.org/10.37128/2306-8744-2019-4-6.

Full text
Abstract:
The work is devoted to the study of the influence of the initial velocity of the loose mixture on the loading process of the vibrating sieve. The regularities of layer thickness, longitudinal and transverse components of velocity, density of loose mixture and specific load on the entire area of a vibrating sieve are established. When the initial velocity is less than velocity of the mixture movement on the sieve is the thickness of the layer has become over the entire surface area, the surface density of the mixture decreases, and the longitudinal velocity component increases with length. The transverse velocity component contributes to the rapid redistribution of the mixture from the congested central area to the unloaded lateral ones. When the initial velocity is equal to the velocity of the mixture movement on the sieve, the thickness of the layer and the surface density of the mixture are constant on the surface area, the longitudinal velocity component is constant along the length and has an initial velocity profile along the width of the sieve, which is aligned with the length. The transverse velocity component decreased and the specific loading deviations increased. When the initial velocity is greater than the velocity of the mixture movement on the sieve, the thickness of the layer decreases, the surface density of the mixture increases, and the longitudinal velocity component decreases with length. The transverse velocity component is almost absent, the specific loading is uneven throughout the sieve area. Thus, the value of the initial velocity affects all the characteristics of the loose mixture, and the nature of changing some of them turns to the opposite. When the mixture is unevenly fed across the width at the inlet of the sieve, the increase of the initial velocity increases the uneven distribution of the specific load over the area of the work surface. The regularities of distribution of the specific load of the sieve are decisive in the design of feeders and distributors of loose mixtures, as well as in calculation of separation modes.
APA, Harvard, Vancouver, ISO, and other styles
39

Banik, N. C. "An effective anisotropy parameter in transversely isotropic media." GEOPHYSICS 52, no. 12 (December 1987): 1654–64. http://dx.doi.org/10.1190/1.1442282.

Full text
Abstract:
An interesting physical meaning is presented for the anisotropy parameter δ, previously introduced by Thomsen to describe weak anisotropy in transversely isotropic media. Roughly, δ is the difference between the P-wave and SV-wave anisotropies of the medium. The observed systematic depth errors in the North Sea are reexamined in view of the new interpretation of the moveout velocity through δ. The changes in δ at an interface adequately describe the effects of transverse isotropy on the P-wave reflection amplitude, The reflection coefficient expression is linearized in terms of changes in elastic parameters. The linearized expression clearly shows that it is the variation of δ at the interface that gives the anisotropic effects at small incidence angles. Thus, δ effectively describes both the moveout velocity and the reflection amplitude variation, two very important pieces of information in reflection seismic prospecting, in the presence of transverse isotropy.
APA, Harvard, Vancouver, ISO, and other styles
40

Scherrer, Robert J., and Abraham Loeb. "The relation between transverse and radial velocity distributions for observations of an isotropic velocity field." Monthly Notices of the Royal Astronomical Society: Letters 483, no. 1 (December 27, 2018): L132—L137. http://dx.doi.org/10.1093/mnrasl/sly232.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Zhou, Yuanlong, Haiquan Bi, Honglin Wang, and Bo Lei. "Critical velocity in the transverse passages of a railway tunnel rescue station with semi-transverse ventilation." Tunnelling and Underground Space Technology 92 (October 2019): 103064. http://dx.doi.org/10.1016/j.tust.2019.103064.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Wei, Xiaohui, Chunlei Wang, Yongjun Long, and Shigang Wang. "Pitch angular velocity dynamics property in transverse galloping pattern from the passive dynamic perspective." Proceedings of the Institution of Mechanical Engineers, Part C: Journal of Mechanical Engineering Science 229, no. 6 (July 15, 2014): 1060–73. http://dx.doi.org/10.1177/0954406214543293.

Full text
Abstract:
This paper employs the simplified quadrupedal passive dynamic model to analyze the underlying property of the transverse gallop. First, the simplified sagittal quadrupedal planar model of the transverse gallop, the trot, and searching methods for achieving the periodic motion are introduced. Next, we explore the dynamic performance of the transverse galloping pattern on the simplified quadrupedal passive dynamic model at different forward speeds and different stiffnesses of legs in a range of initial angular velocities at which the periodic motion is achieved, and there exists a particular property that is defined as the pitch angular velocity dynamics (PAVD) property. If the forward speed and the stiffness of legs are adequate in the transverse galloping pattern, the PAVD emerges and the forces (e.g. peak ground reaction force and average vertical force) that the model undergoes during locomotion are sharply reduced. The Froude numbers of the minimum forward speed where PAVD holds at a fixed stiffness of legs for different samples are virtually the same. Moreover, a comparison between the trot and the transverse gallop illustrates that a trot is more favorable in the range of forward speeds where the PAVD property is not achieved, while a transverse gallop is better after the property emerges.
APA, Harvard, Vancouver, ISO, and other styles
43

Tsvankin, Ilya. "Reflection moveout and parameter estimation for horizontal transverse isotropy." GEOPHYSICS 62, no. 2 (March 1997): 614–29. http://dx.doi.org/10.1190/1.1444170.

Full text
Abstract:
Transverse isotropy with a horizontal axis of symmetry (HTI) is the simplest azimuthally anisotropic model used to describe fractured reservoirs that contain parallel vertical cracks. Here, I present an exact equation for normal‐moveout (NMO) velocities from horizontal reflectors valid for pure modes in HTI media with any strength of anisotropy. The azimuthally dependent P‐wave NMO velocity, which can be obtained from 3-D surveys, is controlled by the principal direction of the anisotropy (crack orientation), the P‐wave vertical velocity, and an effective anisotropic parameter equivalent to Thomsen's coefficient δ. An important parameter of fracture systems that can be constrained by seismic data is the crack density, which is usually estimated through the shear‐wave splitting coefficient γ. The formalism developed here makes it possible to obtain the shear‐wave splitting parameter using the NMO velocities of P and shear waves from horizontal reflectors. Furthermore, γ can be estimated just from the P‐wave NMO velocity in the special case of the vanishing parameter ε, corresponding to thin cracks and negligible equant porosity. Also, P‐wave moveout alone is sufficient to constrain γ if either dipping events are available or the velocity in the symmetry direction is known. Determination of the splitting parameter from P‐wave data requires, however, an estimate of the ratio of the P‐to‐S vertical velocities (either of the split shear waves can be used). Velocities and polarizations in the vertical symmetry plane of HTI media, that contains the symmetry axis, are described by the known equations for vertical transverse isotropy (VTI). Time‐related 2-D P‐wave processing (NMO, DMO, time migration) in this plane is governed by the same two parameters (the NMO velocity from a horizontal reflector and coefficient ε) as in media with a vertical symmetry axis. The analogy between vertical and horizontal transverse isotropy makes it possible to introduce Thomsen parameters of the “equivalent” VTI model, which not only control the azimuthally dependent NMO velocity, but also can be used to reconstruct phase velocity and carry out seismic processing in off‐symmetry planes.
APA, Harvard, Vancouver, ISO, and other styles
44

Eckels, P. W., and T. J. Rabas. "Dehumidification: On the Correlation of Wet and Dry Transport Processes in Plate Finned-Tube Heat Exchangers." Journal of Heat Transfer 109, no. 3 (August 1, 1987): 575–82. http://dx.doi.org/10.1115/1.3248127.

Full text
Abstract:
The transverse velocity of the condensing phase during dehumidification is analogous to the transverse velocity at the wall when exercising boundary layer control by fluid extraction through a permeable wall. Wet and dry pressure drop and heat transfer rates are analyzed for correlation using boundary layer suction theory. Data are presented for flat-plate finned-tube heat exchangers during air heating and dehumidification operations and the data show a significant effect of transverse velocity correlated by the boundary layer suction formulation. The condensate film is considered isothermal in this analysis and the results indicate that an improved modeling of the condensate film is required. We find that the transverse velocity of the condensing phase has an important effect on transport phenomena during dehumidification and that the validity of the Chilton–Colburn heat and mass transfer analogy in describing dehumidification is supported by these results. It should be noted that the dry data form the beginning of a plate fin heat exchanger data base. The present data show the effect of tube diameter and, independently, fin density variation on the Colburn and friction factors with all other geometric parameters held invariant.
APA, Harvard, Vancouver, ISO, and other styles
45

Deller, A. T., J. M. Weisberg, D. J. Nice, and S. Chatterjee. "A VLBI Distance and Transverse Velocity for PSR B1913+16." Astrophysical Journal 862, no. 2 (August 1, 2018): 139. http://dx.doi.org/10.3847/1538-4357/aacf95.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Melekhova, N. A. "Mechanotron transducer of transverse velocity fluctuations of the oncoming flow." Hydrotechnical Construction 23, no. 7 (July 1989): 415–17. http://dx.doi.org/10.1007/bf01431872.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

NOULLEZ, A., G. WALLACE, W. LEMPERT, R. B. MILES, and U. FRISCH. "Transverse velocity increments in turbulent flow using the RELIEF technique." Journal of Fluid Mechanics 339 (May 25, 1997): 287–307. http://dx.doi.org/10.1017/s0022112097005338.

Full text
Abstract:
Non-intrusive measurements of the streamwise velocity in turbulent round jets in air are performed by recording short-time displacements and distorsions of very thin tagging lines written spanwise into the flow. The lines are written by Raman-exciting oxygen molecules and are interrogated by laser-induced electronic fluorescence (relief). This gives access to the spatial structure of transverse velocity increments without recourse to the Taylor hypothesis. The resolution is around 25 μm, less than twice the Kolmogorov scale η for the experiments performed (with Rλ≈360–600).The technique is validated by comparison with results obtained from other techniques for longitudinal or transverse structure functions up to order 8. The agreement is consistent with the estimated errors – a few percent on exponents determined by extended-self-similarity – and indicates significant departures from Kolmogorov (1941) scaling.Probability distribution functions of transverse velocity increments Δu over separations down to 1:8η are reported for the first time. Violent events, with Δu comparable to the r.m.s. turbulent velocity fluctuation, are found to take place with statistically significant probabilities. The shapes of the corresponding lines suggest the effect of intense slender vortex filaments.
APA, Harvard, Vancouver, ISO, and other styles
48

Shen, Zhiyuan, Nan Zeng, and Yonghong He. "Transverse flow-velocity quantification using optical coherence tomography with correlation." Journal of Physics: Conference Series 277 (January 1, 2011): 012033. http://dx.doi.org/10.1088/1742-6596/277/1/012033.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Wang, Yi, and Ruikang Wang. "Autocorrelation optical coherence tomography for mapping transverse particle-flow velocity." Optics Letters 35, no. 21 (October 18, 2010): 3538. http://dx.doi.org/10.1364/ol.35.003538.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Milic, D., M. D. Hoogerland, K. G. H. Baldwin, and R. E. Scholten. "Transverse laser cooling of a velocity-selected sodium atomic beam." Quantum and Semiclassical Optics: Journal of the European Optical Society Part B 8, no. 3 (June 1996): 629–40. http://dx.doi.org/10.1088/1355-5111/8/3/023.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography