Journal articles on the topic 'Plane Wake'

To see the other types of publications on this topic, follow the link: Plane Wake.

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 'Plane Wake.'

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

ROGERS, MICHAEL M. "The evolution of strained turbulent plane wakes." Journal of Fluid Mechanics 463 (July 25, 2002): 53–120. http://dx.doi.org/10.1017/s0022112002008686.

Full text
Abstract:
Direct numerical simulations of ten turbulent time-evolving strained wakes have been generated using a pseudo-spectral numerical method. In all the simulations, the strain was applied to the same (previously generated) initial developed self-similar wake flow field. The cases include flows in which the wake is subjected to various orientations of the applied mean strain, including both plane and axisymmetric strain configurations. In addition, for one particular strain geometry, cases with differing strain rates were considered. Although classical self-similar analysis does yield a self-similar solution for strained wakes, this solution does not describe the observed flow evolution. Instead, the wake mean velocity profiles evolve according to a different ‘equilibrium similarity solution’, with the strained wake width being determined by the straining in the inhomogeneous cross-stream direction. Wakes that are compressed in this direction eventually exhibit constant widths, whereas wakes in cases with expansive cross-stream strain ultimately spread at the same rate as the distortion caused by the applied strain. The shape of the wake mean velocity deficit profile is nearly universal. Although the effect of the strain on the mean flow is pronounced and rapid, the response of the turbulence to the strain occurs more slowly. Changes in the turbulence intensity cannot keep pace with changes in the mean wake velocity deficit, even for relatively low strain rates.
APA, Harvard, Vancouver, ISO, and other styles
2

Weygandt, James H., and Rabindra D. Mehta. "Three-dimensional structure of straight and curved plane wakes." Journal of Fluid Mechanics 282 (January 10, 1995): 279–311. http://dx.doi.org/10.1017/s0022112095000140.

Full text
Abstract:
The formation and evolution of the three-dimensional structure of straight and mildly curved ($b/\bar{R} < 2\%$) flat plate wakes at relatively high Reynolds numbers (Reb = 28 000) have been studied through detailed measurements of the mean and fluctuating velocities. In both cases, the role of initial conditions was examined by generating wakes from untripped (laminar) and tripped (turbulent) initial boundary layers. The curved wake was affected by the angular momentum instability such that the inside half of the wake was unstable, whereas the outside half was stable. In both the straight and curved untripped wakes, large spanwise variations, in the form of ‘pinches’ and ‘crests’, were observed in the contours of mean velocity and Reynolds stresses. Well-organized, ‘spatially stationary’ streamwise vorticity was generated in the near-field region in the form of quadrupoles, to which the spanwise variations in the velocity contours were attributed. The presence of mean streamwise vorticity had a significant effect on the wake growth and defect decay rates, mainly by providing additional entrainment. In the straight wake, the mean streamwise vorticity decayed on both sides of the wake such that it had decayed completely by the far-field region. However, in the curved case, the mean streamwise vorticity on the unstable side decayed at a rate significantly lower than that on the stable side. Despite the decay of mean streamwise vorticity, the spanwise variations persisted into the far wake in both cases. The effects of curvature were also apparent in the Reynolds stress results which showed that the levels on the unstable side were increased significantly compared to those on the stable side, with the effect much stronger in the initially laminar wake. With the initial boundary layers tripped, spatially stationary streamwise vortex structures were not observed in either the straight or curved wakes and the velocity contours appeared nominally two-dimensional. This result further confirms the strong dependency of the three-dimensional structure of plane wakes on initial conditions.
APA, Harvard, Vancouver, ISO, and other styles
3

TAMMISOLA, OUTI, FREDRIK LUNDELL, PHILIPP SCHLATTER, ARMIN WEHRFRITZ, and L. DANIEL SÖDERBERG. "Global linear and nonlinear stability of viscous confined plane wakes with co-flow." Journal of Fluid Mechanics 675 (April 4, 2011): 397–434. http://dx.doi.org/10.1017/jfm.2011.24.

Full text
Abstract:
The global stability of confined uniform density wakes is studied numerically, using two-dimensional linear global modes and nonlinear direct numerical simulations. The wake inflow velocity is varied between different amounts of co-flow (base bleed). In accordance with previous studies, we find that the frequencies of both the most unstable linear and the saturated nonlinear global mode increase with confinement. For wake Reynolds numberRe= 100 we find the confinement to be stabilising, decreasing the growth rate of the linear and the saturation amplitude of the nonlinear modes. The dampening effect is connected to the streamwise development of the base flow, and decreases for more parallel flows at higherRe. The linear analysis reveals that the critical wake velocities are almost identical for unconfined and confined wakes atRe≈ 400. Further, the results are compared with literature data for an inviscid parallel wake. The confined wake is found to be more stable than its inviscid counterpart, whereas the unconfined wake is more unstable than the inviscid wake. The main reason for both is the base flow development. A detailed comparison of the linear and nonlinear results reveals that the most unstable linear global mode gives in all cases an excellent prediction of the initial nonlinear behaviour and therefore the stability boundary. However, the nonlinear saturated state is different, mainly for higherRe. ForRe= 100, the saturated frequency differs less than 5% from the linear frequency, and trends regarding confinement observed in the linear analysis are confirmed.
APA, Harvard, Vancouver, ISO, and other styles
4

CHEN, DAOYI, and GERHARD H. JIRKA. "Absolute and convective instabilities of plane turbulent wakes in a shallow water layer." Journal of Fluid Mechanics 338 (May 10, 1997): 157–72. http://dx.doi.org/10.1017/s0022112097005041.

Full text
Abstract:
In shallow turbulent wake flows (typically an island wake), the flow patterns have been found experimentally to depend mainly on a shallow wake parameter, S=cfD/h in which cf is a quadratic-law friction coefficient, D is the island diameter and h is water depth. In order to understand the dependence of flow patterns on S, the shallow-water stability equation (a modified Orr–Sommerfeld equation) has been derived from the depth-averaged equations of motion with terms which describe bottom friction. Absolute and convective instabilities have been investigated on the basis of wake velocity profiles with a velocity deficit parameter R. Numerical computations have been carried out for a range of R-values and a stability diagram with two dividing lines was obtained, one defining the boundary between absolute and convective instabilities Sca, and another defining the transition between convectively unstable and stable wake flow Scc. The experimental measurements (Chen & Jirka 1995) of return velocities in shallow wakes were used to compute R-values and two critical values, SA=0.79 and SC=0.85, were obtained at the intersections with lines Sca and Scc. Through comparison with transition values observed experimentally for wakes with unsteady bubble (recirculation zone) and vortex shedding, SU and SV respectively, the sequence SC>SA> SU>SV shows vortex shedding to be the end product of absolute instability. This is analogous to the sequence of critical Reynolds numbers for an unbounded wake of large spanwise extent. Experimental frequency characteristics compare well with theoretical results. The observed values of SU and SV for different flow patterns correspond to the velocity profile with R=−0.945, which is located at the end of the wake bubble, and it provides the dominant mode.
APA, Harvard, Vancouver, ISO, and other styles
5

EWING, D., W. K. GEORGE, M. M. ROGERS, and R. D. MOSER. "Two-point similarity in temporally evolving plane wakes." Journal of Fluid Mechanics 577 (April 19, 2007): 287–307. http://dx.doi.org/10.1017/s0022112006003260.

Full text
Abstract:
The governing equations for the two-point correlations of the turbulent fluctuating velocity in the temporally evolving wake were analysed to determine whether they could have equilibrium similarity solutions. It was found that these equations could have such solutions for a finite-Reynolds-number wake, where the two-point velocity correlations could be written as a product of a time-dependent scale and a function dependent only on similarity variables. It is therefore possible to collapse the two-point measures of all the scales of motions in the temporally evolving wake using a single set of similarity variables. As in an earlier single-point analysis, it was found that the governing equations for the equilibrium similarity solutions could not be reduced to a form that was independent of a growth-rate dependent parameter. Thus, there is not a single ‘universal’ solution that describes the state of the large-scale structures, so that the large-scale structures in the far field may depend on how the flow is generated.The predictions of the similarity analysis were compared to the data from two direct numerical simulations of the temporally evolving wakes examined previously. It was found that the two-point velocity spectra of these temporally evolving wakes collapsed reasonably well over the entire range of scales when they were scaled in the manner deduced from the equilibrium similarity analysis. Thus, actual flows do seem to evolve in a manner consistent with the equilibrium similarity solutions.
APA, Harvard, Vancouver, ISO, and other styles
6

Neu, W., P. Mitra, and J. Schetz. "The Wake of Self-Propelled and Over-Thrusted Slender Bodies Near a Simulated Free Surface." Journal of Ship Research 32, no. 01 (March 1, 1988): 70–79. http://dx.doi.org/10.5957/jsr.1988.32.1.70.

Full text
Abstract:
Measurements were performed in the turbulent wake of a propeller-driven axisymmetric body with a plane of symmetry. A flat plate strut was attached to the upper surface of the axisymmetric body, giving a configuration like that of a SWATH-type ship, with the free surface replaced by the plane of symmetry. All mean flow and turbulent flow parameters were measured at three streamwise stations. The measurements were performed for the self-propelled condition and 100 percent over-thrust condition. In the far wake, the center of the wake was found to migrate towards the plane of symmetry. Some interactions were noted between the wakes of the propeller-driven axisymmetric body and that of the flat plate strut—yielding lower axial velocities, higher turbulence intensities and larger static pressure changes compared to regions free of such interference. Comparisons of these effects in the self-propelled case, 100 percent over-thrust case and a previous unpropelled case are given. Spectral measurements were also performed in both near-wake and far-wake regions.
APA, Harvard, Vancouver, ISO, and other styles
7

Kraft, Wayne N., and Malcolm J. Andrews. "Experimental Investigation of Unstably Stratified Buoyant Wakes." Journal of Fluids Engineering 128, no. 3 (November 1, 2005): 488–93. http://dx.doi.org/10.1115/1.2174060.

Full text
Abstract:
A water channel has been used as a statistically steady experiment to investigate the development of a buoyant plane wake. Parallel streams of hot and cold water are initially separated by a splitter plate and are oriented to create an unstable stratification. At the end of the splitter plate, the two streams are allowed to mix and a buoyancy-driven mixing layer develops. The continuous, unstable stratification inside the developing mixing layer provides the necessary environment to study the buoyant wake. Downstream a cylinder was placed at the center of the mixing layer. As a result the dynamic flows of the plane wake and buoyancy-driven mixing layer interact. Particle image velocimetry and a high-resolution thermocouple system have been used to measure the response of the plane wake to buoyancy driven turbulence. Velocity and density measurements are used as a basis from which we describe the transition, and return to equilibrium, of the buoyancy-driven mixing layer. Visual observation of the wake does not show the usual vortex street associated with a cylinder wake, but the effect of the wake is apparent in the measured vertical velocity fluctuations. An expected peak in velocity fluctuations in the wake is found, however the decay of vertical velocity fluctuations occurs at a reduced rate due to vertical momentum transport into the wake region from buoyancy-driven turbulence. Therefore for wakes where buoyancy is driving the motion, a remarkably fast recovery of a buoyancy-driven Rayleigh-Taylor mixing in the wake region is found.
APA, Harvard, Vancouver, ISO, and other styles
8

Key, Nicole L. "Influence of Upstream and Downstream Compressor Stators on Rotor Exit Flow Field." International Journal of Rotating Machinery 2014 (2014): 1–10. http://dx.doi.org/10.1155/2014/392352.

Full text
Abstract:
Measurements acquired at the rotor exit plane illuminate the interaction of the rotor with the upstream vane row and the downstream vane row. The relative phase of the upstream and downstream vane rows is adjusted using vane clocking so that the effect of the upstream propagating potential field from the downstream stator can be distinguished from the effects associated with the wakes shed from the upstream stator. Unsteady absolute flow angle information shows that the downstream potential field causes the absolute flow angle to increase in the vicinity of the downstream stator leading edge. The presence of Stator 1 wake is also detected at this measurement plane using unsteady total pressure data. The rotor wakes are measured at different circumferential locations across the vane passage, and the influence of Stator 1 wake on the suction side of the rotor wake is evident. Also, the influence of the downstream stator is detected on the pressure side of the rotor wake for a particular clocking configuration. Understanding the role of the surrounding vane rows on rotor wake development will lead to improved comparison between experimental data and results from computational models.
APA, Harvard, Vancouver, ISO, and other styles
9

Zhou, Y., and R. A. Antonia. "Critical points in a turbulent near wake." Journal of Fluid Mechanics 275 (September 25, 1994): 59–81. http://dx.doi.org/10.1017/s0022112094002284.

Full text
Abstract:
Velocity data were obtained in the turbulent wake of a circular cylinder with an orthogonal array of sixteen X-wires, eight in the (x, y)-plane and eight in the (x, z)-plane. By applying the phase-plane technique to these data, three types of critical points (where the velocity is zero and the streamline slope is indeterminate) were identified. Of these, foci and saddle points occurred most frequently, although a significant number of nodes was also found. Flow topology and properties associated with these points were obtained in each plane. Saddle-point regions associated with spanwise vortices provide the dominant contribution to the Reynolds shear stress and larger contributions to the normal stresses than focal regions. The topology was found to be in close agreement with that obtained from other methods of detecting features of the organized motion. The inter-relationship between critical points simultaneously identified in the two planes can provide some insight into the three-dimensionality of the organized motion. Foci in the (x, z)-plane correspond, with relatively high probability and almost negligible streamwise separation, to saddle points in the (x, y)-plane and are interpreted in terms of ribs aligned with the diverging separatrix between consecutive spanwise vortex rolls. Foci in the (x, z)-plane which correspond, with relatively weak probability, to foci in the (x, y)-plane seem consistent with a distortion of the vortex rolls in the (y, z)-plane.
APA, Harvard, Vancouver, ISO, and other styles
10

El Khoury, George K., Helge I. Andersson, and Bjørnar Pettersen. "Wakes behind a prolate spheroid in crossflow." Journal of Fluid Mechanics 701 (May 18, 2012): 98–136. http://dx.doi.org/10.1017/jfm.2012.135.

Full text
Abstract:
AbstractViscous laminar flow past a prolate $(L/ d= 6)$ spheroid has been investigated numerically at seven different Reynolds numbers; $\mathit{Re}= 50, 75, 100, 150, 200, 250$ and $300$. In contrast to all earlier investigations, the major axis of the spheroid was oriented perpendicular to the free stream flow. As expected, the flow field in the wake showed a strong resemblance of that observed behind a finite-length circular cylinder, yet had features observed in the axisymmetric wake behind a sphere. The following different flow regimes were observed in the present computational study: (i) steady laminar flow with massive flow separation and symmetry about the equatorial and the meridional planes at $\mathit{Re}= 50$; (ii) steady laminar flow with massive flow separation and symmetry about the equatorial and the meridional plane at $\mathit{Re}= 75$, but the flow in the equatorial plane did no longer resemble the steady wake behind a circular cylinder; (iii) unsteady laminar flow with Strouhal number $0. 109$ and symmetry about the equatorial plane at $\mathit{Re}= 100$; (iv) unsteady laminar flow with two distinct frequencies and without any planar symmetries at $\mathit{Re}= 200$; (v) transitional flow with a dominant shedding frequency $\mathit{St}= 0. 151$ and without any spatial symmetries at $\mathit{Re}= 300$. For all but the two lowest $\mathit{Re}$ hairpin vortices were alternately shed from the two sides of the spheroid and resulted in a ladder-like pattern of oppositely oriented vortex structures, in contrast with the single-sided shedding in the wake of a sphere. The contour of the very-near-wake mimicked the shape of the prolate spheroid. However, $15d$ downstream the major axis of the wake became aligned with the minor axis of the spheroid. This implies that an axis switching occurred some $10d$ downstream, i.e. the cross-section of the wake evolved such that the major and minor axes interchanged at a certain downstream location. This peculiar phenomenon has frequently been reported to arise for elliptical and rectangular jets, whereas observations of axis switching for asymmetric wakes are scarce.
APA, Harvard, Vancouver, ISO, and other styles
11

Hayakawa, Michio, and Fazle Hussain. "Three-dimensionality of organized structures in a plane turbulent wake." Journal of Fluid Mechanics 206 (September 1989): 375–404. http://dx.doi.org/10.1017/s0022112089002338.

Full text
Abstract:
This paper describes a quantitative study of the three-dimensional nature of organized motions in a turbulent plane wake. Coherent structures are detected from the instantaneous, spatially phase-correlated vorticity field using certain criteria based on size, strength and geometry of vortical structures. With several combinations of X-wire rakes, vorticity distributions in the spanwise and transverse planes are measured in the intermediate region (10d [les ] x [les ] 40d) of the plane turbulent wake of a circular cylinder at a Reynolds number of 13000 based on the cylinder diameter d. Spatial correlations of smoothed vorticity signals as well as phase-aligned ensemble-averaged vorticity maps over structure cross-sections yield a quantitative measure of the spatial coherence and geometry of organized structures in the fully turbulent field. The data demonstrate that the organized structures in the nominally two-dimensional wake exhibit significant three-dimensionality even in the near field. Using instantaneous velocity and vorticity maps as well as correlations of vorticity distributions in different planes, some topological features of the dominant coherent structures in a plane wake are inferred.
APA, Harvard, Vancouver, ISO, and other styles
12

MELIGA, PHILIPPE, DENIS SIPP, and JEAN-MARC CHOMAZ. "Absolute instability in axisymmetric wakes: compressible and density variation effects." Journal of Fluid Mechanics 600 (March 26, 2008): 373–401. http://dx.doi.org/10.1017/s0022112008000499.

Full text
Abstract:
Lesshafft & Huerre (Phys. Fluids, 2007; vol. 19, 024102) have recently studied the transition from convective to absolute instability in hot round jets, for which absolute instability is led by axisymmetric perturbations and enhanced when lowering the jet density. The present paper analyses similarly the counterpart problem of wake flows, and establishes that absolute instability is then led by a large-scale helical wake mode favoured when the wake is denser than the surrounding fluid. This generalizes to variable density and compressible wakes the results of Monkewitz (J. Fluid Mech. vol 192, 1988, p. 561). Furthermore, we show that in a particular range of density ratios, the large-scale helical wake mode can become absolutely unstable by increasing only the Mach number up to high subsonic values. This possibility of an absolute instability triggered by an increase of the Mach number is opposite to the behaviour previously described in shear flows such as plane mixing layers and axisymmetric jets. A physical interpretation based on the action of the baroclinic torque is proposed. An axisymmetric short-scale mode, similar to that observed in plane mixing layers, leads the transition in light wakes, but the corresponding configurations require large counterflow for the instability to be absolute.These results suggest that the low-frequency oscillation present in afterbody wakes may be due to a nonlinear global mode triggered by a local absolute instability, since the azimuthal wavenumber and absolute frequency of the helical wake mode agree qualitatively with observations.
APA, Harvard, Vancouver, ISO, and other styles
13

Antonia, R. A., and L. W. B. Browne. "Heat transport in a turbulent plane wake." International Journal of Heat and Mass Transfer 29, no. 10 (October 1986): 1585–92. http://dx.doi.org/10.1016/0017-9310(86)90073-6.

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

Zhou, Y., and R. A. Antonia. "Memory effects in a turbulent plane wake." Experiments in Fluids 19, no. 2 (June 1995): 112–20. http://dx.doi.org/10.1007/bf00193857.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Dmitrenko, Yu M., I. I. Kovalev, N. N. Luchko, and P. Ya Cherepanov. "Plane turbulent wake with zero excess momentum." Journal of Engineering Physics 52, no. 5 (May 1987): 536–42. http://dx.doi.org/10.1007/bf00873306.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

van der Deijl, W., M. Obligado, C. Sicot, and S. Barre. "Experimental study of mean and turbulent velocity fields in the wake of a twin-rotor vertical axis wind turbine." Journal of Physics: Conference Series 2265, no. 2 (May 1, 2022): 022073. http://dx.doi.org/10.1088/1742-6596/2265/2/022073.

Full text
Abstract:
Abstract The wake of a twin-rotor vertical axis wind turbine (VAWT) was measured experimentally in a wind tunnel using particle image velocimetry. Only the horizontal centre plane was measured, showing that there is no wake recovery in this central plane. The wake deficit in this plane actually increases up to x = 5.5D, independent of tip-speed-ratio. The wake of a VAWT is, therefore, not homogeneous but three-dimensional. Which also means that the aspect ratio of a VAWT is an important parameter if one wants to optimise for wake recovery. In addition, the shed vortices at the edge of the wake do not play a role in the wake recovery. They do not enhance the recovery, nor do they shield the wake from re-energising. A reason for this is that their persistence strongly depends on tip-speed-ratio yet them breaking down has no effect on the wake recovery. Furthermore, the turbulence intensity levels at the edge of the wake are relatively low and often lower than the centre part of the wake. Hence, the dominant process of wake recovery of a VAWT must be happening in the vertical plane, which is in correspondence with previous research.
APA, Harvard, Vancouver, ISO, and other styles
17

Pan, Jun-Hua, Nian-Mei Zhang, and Ming-Jiu Ni. "The wake structure and transition process of a flow past a sphere affected by a streamwise magnetic field." Journal of Fluid Mechanics 842 (March 7, 2018): 248–72. http://dx.doi.org/10.1017/jfm.2018.133.

Full text
Abstract:
The wake structure and transition process of an incompressible viscous fluid flow past a sphere affected by an imposed streamwise magnetic field are investigated numerically over flow regimes that include steady and unsteady laminar flows at Reynolds numbers up to 300. For cases without a magnetic field, a subregion with the existence of a limit cycle is found in the range $210<Re<270$. The point of division is between $Re=220$ and $Re=230$. For cases with a streamwise magnetic field, five wake patterns are the steady axisymmetric wake with an attached separation bubble, the steady plane symmetric wake with a small spiral dismissed, the steady plane symmetric wake with a limit cycle, the steady plane symmetric wake with a small spiral fed by the upstream fluid and the unsteady plane symmetric wake with a wave-like oscillation or vortex shedding. Under the influence of an imposed streamwise magnetic field, the wake will be transitioned to various patterns. An interesting ‘reversion phenomenon’, which describes the topological structure behind a sphere with a higher Reynolds number and a certain interaction parameter which corresponds to a lower Reynolds number case with a certain interaction parameter or a much lower Reynolds number case without a magnetic field, is also found. The principal results of the present work are summarized in a map of regimes in the $\{N,Re\}$ plane.
APA, Harvard, Vancouver, ISO, and other styles
18

Farrell, W., T. Herges, D. Maniaci, and K. Brown. "Wake state estimation of downwind turbines using recurrent neural networks for inverse dynamics modelling." Journal of Physics: Conference Series 2265, no. 3 (May 1, 2022): 032094. http://dx.doi.org/10.1088/1742-6596/2265/3/032094.

Full text
Abstract:
Abstract Presented in this work is a novel approach to estimate absolute lateral wake center position on the rotor plane of a waked turbine using turbine load and operating state information. The approach formulates the estimation of the absolute lateral wake position as an inverse dynamics problem and utilizes a recurrent neural network to model the inverse mapping between the wake center position and select turbine output channels. The technique is validated on experimental data collected from experiments at the Scaled Wind Farm Technology (SWiFT) facility and numerical simulations of the site in the wind farm simulator FAST.Farm. Estimator performance and analysis of optimal conditions for estimation are discussed.
APA, Harvard, Vancouver, ISO, and other styles
19

Kopp, G. A., J. G. Kawall, and J. F. Keffer. "The evolution of the coherent structures in a uniformly distorted plane turbulent wake." Journal of Fluid Mechanics 291 (May 25, 1995): 299–322. http://dx.doi.org/10.1017/s0022112095002710.

Full text
Abstract:
A plane turbulent wake generated by a flat plate is subjected to a uniform distortion. It is observed that nearly two-dimensional, quasi-periodic coherent structures dominate the distorted wake. Rapid distortion theory, applied to a kinematic vortex model of the coherent structures in the undistorted far wake, predicts many of the effects revealed by a hot-wire anemometry/pattern-recognition analysis of these structures. Specifically, rapid distortion theory predicts reasonably well the observed changes in the ensemble-averaged velocity patterns and the disproportionate amplification of the large-scale coherent structures relative to the smaller-scale ‘isotropic’ eddies. These results are consistent with the view that self-preservation of the distorted wake is not possible because of the selective amplification of the coherent structures, which control the development of the wake. As well, the entrainment rate in the distorted wake increases at a rate greater than that predicted by the self-preservation theory.
APA, Harvard, Vancouver, ISO, and other styles
20

Grinstein, F. F., J. P. Boris, and O. M. Griffin. "Passive pressure-drag control in a plane wake." AIAA Journal 29, no. 9 (September 1991): 1436–42. http://dx.doi.org/10.2514/3.10757.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Hajj, M. R., and I. M. Janajreh. "Spectral Energy Transfer in Transition of Plane Wake." Journal of Engineering Mechanics 123, no. 8 (August 1997): 837–42. http://dx.doi.org/10.1061/(asce)0733-9399(1997)123:8(837).

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Yang, Y., C. T. Crowe, J. N. Chung, and T. R. Troutt. "Experiments on particle dispersion in a plane wake." International Journal of Multiphase Flow 26, no. 10 (October 2000): 1583–607. http://dx.doi.org/10.1016/s0301-9322(99)00105-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Selby, G. V., and F. H. Miandoab. "Asymptotic Wake Behavior of Swept, Blunt Trailing-Edge Airfoils." Journal of Fluids Engineering 116, no. 1 (March 1, 1994): 171–74. http://dx.doi.org/10.1115/1.2910232.

Full text
Abstract:
The effect of base sweep and the addition of passive flow-control devices at constant base sweep angle (30 deg) on the asymptotic behavior of turbulent wakes produced by flatplate airfoils was experimentally examined. It was determined that values of the nondimensional streamwise velocity defect and wake thickness parameters for the grooved model with 30 deg swept base at fourteen base thicknesses downstream of the base at mid-span were closer to asymptotic values from empirical plane wake predictions than values for the 0, 30, and 45 deg swept baseline models and the 30 deg swept model with Wishbone vortex generators. The grooves apparently inhibited the three-dimensionality of the resulting wake flow.
APA, Harvard, Vancouver, ISO, and other styles
24

Pan, Jun-Hua, Nian-Mei Zhang, and Ming-Jiu Ni. "Wake structure of laminar flow past a sphere under the influence of a transverse magnetic field." Journal of Fluid Mechanics 873 (June 20, 2019): 151–73. http://dx.doi.org/10.1017/jfm.2019.423.

Full text
Abstract:
The wake structure of an incompressible, conducting, viscous fluid past an electrically insulating sphere affected by a transverse magnetic field is investigated numerically over flow regimes including steady and unsteady laminar flows at Reynolds numbers up to 300. For a steady axisymmetric flow affected by a transverse magnetic field, the wake structure is deemed to be a double plane symmetric state. For a periodic flow, unsteady vortex shedding is first suppressed and transitions to a steady plane symmetric state and then to a double plane symmetric pattern. Wake structures in the range $210<Re\leqslant 300$ without a magnetic field have a symmetry plane. An angle $\unicode[STIX]{x1D703}$ exists between the orientation of this symmetry plane and the imposed transverse magnetic field. For a given transverse magnetic field, the final wake structure is found to be independent of the initial flow configuration with a different angle $\unicode[STIX]{x1D703}$. However, the orientation of the symmetry plane tends to be perpendicular to the magnetic field, which implies that the transverse magnetic field can control the orientation of the wake structure of a free-moving sphere and change the direction of its horizontal motion by a field–wake–trajectory control mechanism. An interesting ‘reversion phenomenon’ is found, where the wake structure of the sphere at a higher Reynolds number and a certain magnetic interaction parameter ($N$) corresponds to a lower Reynolds number with a lower $N$ value. Furthermore, the drag coefficient is proportional to $N^{2/3}$ for weak magnetic fields or to $N^{1/2}$ for strong magnetic fields, where the threshold value between these two regimes is approximately $N=4$.
APA, Harvard, Vancouver, ISO, and other styles
25

Qi, Xiao Ni, Jian Meng, and Yong Qi Liu. "Aerodynamic Analysis of Pickup Truck." Advanced Materials Research 201-203 (February 2011): 1296–99. http://dx.doi.org/10.4028/www.scientific.net/amr.201-203.1296.

Full text
Abstract:
The present study focuses on the aerodynamics of pickup trucks. The CFD software FLUENT was used to simulate flow field around a pickup truck in this paper. Numerical simulation was taken on a 1/5th pickup truck model. The surface pressure distribution, the wake velocity distribution of the special profiles and the flow structures were obtained. The research indicted that there was a recirculation flow region over the bed for pickup truck. The cab shear layer did not interact directly with the tailgate, flowing above the top of the tailgate. There was a downwash flow in the symmetry plane behind the tailgate with no reverse flow region in the symmetry plane, and the formation of two smaller recirculation flow regions was on both sides of the symmetry plane. Mean flow fields in the near wake of the cab showed a weak pair of counter-rotating vortices behind the cab. In the cross-flow planes behind the tailgate, the mean flow fields show strong counter-rotating vortices behind the tailgate. Instantaneous flow fields in the cross-flow planes of the pickup truck near wake showed compact vortex structures located randomly in space.
APA, Harvard, Vancouver, ISO, and other styles
26

MOSER, ROBERT D., MICHAEL M. ROGERS, and DANIEL W. EWING. "Self-similarity of time-evolving plane wakes." Journal of Fluid Mechanics 367 (July 25, 1998): 255–89. http://dx.doi.org/10.1017/s0022112098001426.

Full text
Abstract:
Direct numerical simulations of three time-developing turbulent plane wakes have been performed. Initial conditions for the simulations were obtained using two realizations of a direct simulation from a turbulent boundary layer at momentum-thickness Reynolds number 670. In addition, extra two-dimensional disturbances were added in two of the cases to mimic two-dimensional forcing. The wakes are allowed to evolve long enough to attain approximate self-similarity, although in the strongly forced case this self-similarity is of short duration. For all three flows, the mass-flux Reynolds number (equivalent to the momentum-thickness Reynolds number in spatially developing wakes) is 2000, which is high enough for a short k−5/3 range to be evident in the streamwise one-dimensional velocity spectra.The spreading rate, turbulence Reynolds number, and turbulence intensities all increase with forcing (by nearly an order of magnitude for the strongly forced case), with experimental data falling between the unforced and weakly forced cases. The simulation results are used in conjunction with a self-similar analysis of the Reynolds stress equations to develop scalings that approximately collapse the profiles from different wakes. Factors containing the wake spreading rate are required to bring profiles from different wakes into agreement. Part of the difference between the various cases is due to the increased level of spanwise-coherent (roughly two-dimensional) energy in the forced cases. Forcing also has a significant impact on flow structure, with the forced flows exhibiting more organized large-scale structures similar to those observed in transitional wakes.
APA, Harvard, Vancouver, ISO, and other styles
27

Stallard, T., R. Collings, T. Feng, and J. Whelan. "Interactions between tidal turbine wakes: experimental study of a group of three-bladed rotors." Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 371, no. 1985 (February 28, 2013): 20120159. http://dx.doi.org/10.1098/rsta.2012.0159.

Full text
Abstract:
It is well known that a wake will develop downstream of a tidal stream turbine owing to extraction of axial momentum across the rotor plane. To select a suitable layout for an array of horizontal axis tidal stream turbines, it is important to understand the extent and structure of the wakes of each turbine. Studies of wind turbines and isolated tidal stream turbines have shown that the velocity reduction in the wake of a single device is a function of the rotor operating state (specifically thrust), and that the rate of recovery of wake velocity is dependent on mixing between the wake and the surrounding flow. For an unbounded flow, the velocity of the surrounding flow is similar to that of the incident flow. However, the velocity of the surrounding flow will be increased by the presence of bounding surfaces formed by the bed and free surface, and by the wake of adjacent devices. This paper presents the results of an experimental study investigating the influence of such bounding surfaces on the structure of the wake of tidal stream turbines.
APA, Harvard, Vancouver, ISO, and other styles
28

Yu, W. S., and B. Lakshminarayana. "Numerical Simulation of the Effects of Rotor-Stator Spacing and Wake/Blade Count Ratio on Turbomachinery Unsteady Flows." Journal of Fluids Engineering 117, no. 4 (December 1, 1995): 639–46. http://dx.doi.org/10.1115/1.2817316.

Full text
Abstract:
A two-dimensional time-accurate Navier-Stokes solver for incompressible flows is used to simulate the effects of the axial spacing between an upstream rotor and a stator, and the wake/blade count ratio on turbomachinery unsteady flows. The code uses a pressure-based method. A low-Reynolds number two-equation turbulence model is incorporated to account for the turbulence effect. By computing cases with different spacing between an upstream rotor wake and a stator, the effect of the spacing is simulated. Wake/blade count ratio effect is simulated by varying the number of rotor wakes in one stator passage at the computational inlet plane. Results on surface pressure, unsteady velocity vectors, blade boundary layer profiles, rotor wake decay and loss coefficient for all the cases are interpreted. It is found that the unsteadiness in the stator blade passage increases with a decrease in the blade row spacing and a decrease in the wake/blade count ratio. The reduced frequency effect is dominant in the wake/blade count ratio simulation. The time averaged loss coefficient increases with a decrease in the axial blade row spacing and an increase in the wake/blade count ratio.
APA, Harvard, Vancouver, ISO, and other styles
29

Menke, Robert, Nikola Vasiljević, Kurt S. Hansen, Andrea N. Hahmann, and Jakob Mann. "Does the wind turbine wake follow the topography? A multi-lidar study in complex terrain." Wind Energy Science 3, no. 2 (October 10, 2018): 681–91. http://dx.doi.org/10.5194/wes-3-681-2018.

Full text
Abstract:
Abstract. The wake of a single wind turbine in complex terrain is analysed using measurements from lidars. A particular focus of this analysis is the wake deficit and propagation. Six scanning lidars (three short-range and three long-range WindScanners) were deployed during the Perdigão 2015 measurement campaign, which took place at a double-ridge site in Portugal. Several scanning scenarios, including triple- and dual-Doppler scans, were designed to capture the wind turbine wake of a 2 MW turbine located on one of the ridges. Different wake displacements are categorized according to the time of the day. The results show a strong dependence of the vertical wake propagation on the atmospheric stability. When an atmospheric wave is observed under stable conditions, the wake follows the terrain down the ridge with a maximum inclination of -28∘. During unstable conditions, the wake is advected upwards by up to 29∘ above the horizontal plane.
APA, Harvard, Vancouver, ISO, and other styles
30

EHRHARD, PETER. "Laminar mixed convection in two-dimensional far wakes above heated/cooled bodies: model and experiments." Journal of Fluid Mechanics 439 (July 23, 2001): 165–98. http://dx.doi.org/10.1017/s0022112001004463.

Full text
Abstract:
A heated or cooled body is positioned in a vertically rising forced flow. This develops both a kinematic and a thermal wake, the latter adding buoyant effects to the otherwise forced flow field. An asymptotic model is developed to treat this mixed convection in both plane and axisymmetric geometry. The model holds for laminar flow in the boundary layer approximation and uses a far-wake expansion for weak buoyant forces. For plane geometry the model is validated against both experiments in water and FEM simulations.It is found for a heated wake that buoyant forces accelerate the fluid in the thermal wake such that the vertical velocity deficit in the kinematic wake is reduced. For strong heating this may even lead to vertical velocities larger than the forced flow amplitude. In conjunction the entrainment is intensified in a heated wake. The effects in a cooled wake are opposite in that the vertical velocity deficit is increased within the thermal wake and the horizontal flow into the wake is weakened. In a strongly cooled wake the horizontal flow may even invert, going from the wake centre into the ambient. The Prandtl number controls the width of the thermal wake and, thus, the portion of the kinematic wake which is affected by buoyant forces. Large Prandtl numbers superimpose a narrow buoyant plume, small Prandtl numbers a wide buoyant plume, onto the kinematic wake.
APA, Harvard, Vancouver, ISO, and other styles
31

Bomphrey, Richard J., Per Henningsson, Dirk Michaelis, and David Hollis. "Tomographic particle image velocimetry of desert locust wakes: instantaneous volumes combine to reveal hidden vortex elements and rapid wake deformation." Journal of The Royal Society Interface 9, no. 77 (September 12, 2012): 3378–86. http://dx.doi.org/10.1098/rsif.2012.0418.

Full text
Abstract:
Aerodynamic structures generated by animals in flight are unstable and complex. Recent progress in quantitative flow visualization has advanced our understanding of animal aerodynamics, but measurements have hitherto been limited to flow velocities at a plane through the wake. We applied an emergent, high-speed, volumetric fluid imaging technique (tomographic particle image velocimetry) to examine segments of the wake of desert locusts, capturing fully three-dimensional instantaneous flow fields. We used those flow fields to characterize the aerodynamic footprint in unprecedented detail and revealed previously unseen wake elements that would have gone undetected by two-dimensional or stereo-imaging technology. Vortex iso-surface topographies show the spatio-temporal signature of aerodynamic force generation manifest in the wake of locusts, and expose the extent to which animal wakes can deform, potentially leading to unreliable calculations of lift and thrust when using conventional diagnostic methods. We discuss implications for experimental design and analysis as volumetric flow imaging becomes more widespread.
APA, Harvard, Vancouver, ISO, and other styles
32

Grinstein, F. F., J. P. Boris, and O. M. Griffin. "Errata: Passive-Pressure Drag Control in a Plane Wake." AIAA Journal 29, no. 11 (November 1991): 2016b. http://dx.doi.org/10.2514/3.61528.

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

MAEKAWA, Hiroshi, and Nagi N. MANSOUR. "Direct Numerical Simulations of a Spatially Developing Plane Wake." JSME international journal. Ser. 2, Fluids engineering, heat transfer, power, combustion, thermophysical properties 35, no. 4 (1992): 543–48. http://dx.doi.org/10.1299/jsmeb1988.35.4_543.

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

MOCHIZUKI, Osamu, Hideo YAMADA, Haruo YAMABE, Motomu IMAIZUMI, and Atsuhiro KOBAYASHI. "The wake of the wedge in the plane jet." Transactions of the Japan Society of Mechanical Engineers Series B 51, no. 470 (1985): 3110–18. http://dx.doi.org/10.1299/kikaib.51.3110.

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

MAEKAWA, Hiroshi, and Nagi N. MANSOUR. "Direct Numerical Simulations of a Spatially Developing Plane Wake." Transactions of the Japan Society of Mechanical Engineers Series B 57, no. 540 (1991): 2583–88. http://dx.doi.org/10.1299/kikaib.57.2583.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Tang, L., F. Wen, Y. Yang, C. T. Crowe, J. N. Chung, and T. R. Troutt. "Self‐organizing particle dispersion mechanism in a plane wake." Physics of Fluids A: Fluid Dynamics 4, no. 10 (October 1992): 2244–51. http://dx.doi.org/10.1063/1.858465.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Maekawa, Hiroshi, Nagi N. Mansour, and Jeffrey C. Buell. "Instability mode interactions in a spatially developing plane wake." Journal of Fluid Mechanics 235, no. -1 (February 1992): 223. http://dx.doi.org/10.1017/s0022112092001095.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Davila, J. B., M. R. Hajj, R. W. Miksad, and E. J. Powers. "Wavenumber mismatch of interacting modes in a plane wake." Applied Scientific Research 51, no. 1-2 (June 1993): 385–89. http://dx.doi.org/10.1007/bf01082565.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Pérez-de-Tejada, H., H. Durand-Manterola, M. Reyes-Ruiz, and R. Lundin. "Pluto’s plasma wake oriented away from the ecliptic plane." Icarus 246 (January 2015): 310–16. http://dx.doi.org/10.1016/j.icarus.2014.06.022.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Guvernyuk, S. V., and F. A. Maksimov. "Hysteresis in Supersonic Flow Past a Plane Cascade of Cylindrical Rods." Fluid Dynamics 56, no. 6 (November 2021): 886–96. http://dx.doi.org/10.1134/s0015462821060033.

Full text
Abstract:
Abstract— The results of numerical simulation of the interaction of supersonic flow with a permeable screen in form of an infinite plane cascade (lattice) of circular cylinders are given. The interaction regime in which the shocks ahead of the cylinders are localized on the scale of the cascade step is considered. The multi-block computational technique in which the viscous boundary layers are resolved by means of local grids using the Navier–Stokes equations, while the effects of inteferrence between the shock-wave structures in supersonic wake are described within the framework of Euler’s equations. The action of shock waves induced by the neighboring elements of lattice to the near-wake region behind the intermediate elements can ambiguously affect the aerodynamic lattice performance as well as generate time-dependent phenomena in the wake. The flow regimes are classified depending on continuous increase and decrease in the free-stream supersonic air flow in the Mach number range from 2.4 to 4.2 with reference to the lattice of the 80% permeability. The sources of the hysteresis behavior of the lattice aerodynamic drag with respect to the Mach number and the mechanisms of the onset of self-oscillating wake flow regimes are discussed.
APA, Harvard, Vancouver, ISO, and other styles
41

Mitra, P., W. Neu, and J. Schetz. "Effect of a Simulated Free Surface on the Wake of a Slender Body." Journal of Ship Research 30, no. 04 (December 1, 1986): 242–55. http://dx.doi.org/10.5957/jsr.1986.30.4.242.

Full text
Abstract:
Turbulent flow measurements were performed in the wake of a slender axisymmetric body in the presence of a flat plate strut and an image plane crudely representing the "rigid lid" approximation to a free surface. The tests were performed in a wind tunnel at a nominal Reynolds number of 6.0 ⨯ 105. A Yawhead probe was used for the mean flow measurements, and a Constant Temperature Anemometer System with an x-wire probe was used to obtain the turbulent flow characteristics. The presence of the image plane was found to increase the velocity defect and the static pressure as the image plane was approached. A redistribution among the various components of velocity fluctuations was noted near the image plane. The transverse component was enhanced at the expense of the normal component. The image plane also was found to influence the magnitudes and radial spread of turbulence intensities and Reynolds stresses. Some interactions between the wake of the axisymmetric body and that of the plate strut were observed. Overall, the mean velocities and the turbulence quantities indicated symmetry about the image plane throughout the wake.
APA, Harvard, Vancouver, ISO, and other styles
42

Meiburg, E., and J. C. Lasheras. "Experimental and numerical investigation of the three-dimensional transition in plane wakes." Journal of Fluid Mechanics 190 (May 1988): 1–37. http://dx.doi.org/10.1017/s0022112088001181.

Full text
Abstract:
The three-dimensional structure of a moderate-Reynolds-number (≈ 100) plane wake behind a flat plate subjected to periodic spanwise perturbations has been studied both experimentally and numerically. Comparisons between experimental interface visualizations and numerical calculations demonstrate that important features of the development of the three-dimensional evolution can be reproduced by numerical inviscid vortex dynamics.It is shown that the redistribution, reorientation and stretching of vorticity leads to the formation of counter-rotating pairs of streamwise vortices which superimpose onto the spanwise Kármán-like vortices. These streamwise vortices exhibit lambdashaped structures and are located in the braids connecting consecutive Kármán vortices of opposite sign.The interaction of the evolving streamwise structure with the spanwise Kármán vortices results in the formation of closed vortex loops. Depending on the orientation of the initial perturbation, the three-dimensional vorticity field of the wake acquires either a symmetric or a non-symmetric configuration. Under the effect of a periodic vertical perturbation, the wake develops a non-symmetric vorticity mode exhibiting a staggered array of closed vortex loops of alternating sign. In contrast, under the effect of a periodic horizontal perturbation, the wake acquires a symmetric vorticity mode with the closed vortex loops of alternating sign aligned in the flow direction.
APA, Harvard, Vancouver, ISO, and other styles
43

BURNS, T. J., R. W. DAVIS, and E. F. MOORE. "A perturbation study of particle dynamics in a plane wake flow." Journal of Fluid Mechanics 384 (April 10, 1999): 1–26. http://dx.doi.org/10.1017/s002211209900419x.

Full text
Abstract:
We analyse the dynamics of small, rigid, dilute spherical particles in the far wake of a bluff body under the assumption that the background flow field is approximated by a periodic array of Stuart vortices that can be considered to be a regularization of the von Kármán vortex street. Using geometric singular perturbation theory and numerical methods, we show that when inertia (measured by a dimensionless Stokes number) is not too large, there is a periodic attractor in the phase space of the dynamical system governing the particle motion. We argue that this provides a simple mechanism to explain the unexpected ‘focusing’ effect that has been observed both numerically and experimentally in the far-wake flow past a bluff body by Tang et al. (1992). Their results show that over a range of Reynolds numbers and intermediate values of the Stokes number, particles injected into the wake of a bluff body concentrate near the edges of the vortex structures downstream, thus tending to ‘demix’ rather than disperse homogeneously.
APA, Harvard, Vancouver, ISO, and other styles
44

Hanke, W., C. Brucker, and H. Bleckmann. "The ageing of the low-frequency water disturbances caused by swimming goldfish and its possible relevance to prey detection." Journal of Experimental Biology 203, no. 7 (April 1, 2000): 1193–200. http://dx.doi.org/10.1242/jeb.203.7.1193.

Full text
Abstract:
Wakes caused by swimming goldfish (Carassius auratus) were measured with a particle image velocimetry system and analyzed using a cross-correlation technique. Particle velocities in a horizontal plane (size of measuring plane 24 cmx32 cm or 20 cmx27 cm) were determined, and the vorticity in the plane was derived from these data. The wake behind a swimming goldfish can show a clear vortex structure for at least 30 s. Particle velocities significantly higher than background noise could still be detected 3 min after a fish (body length 10 cm) had passed through the measuring plane. Within this time span, the lateral spread of fish-generated wakes could exceed 30 cm for a 10 cm fish and 20 cm for a 6 cm fish. Measurements in a man-made open-air pond showed that water velocities in a quasi-natural still water environment can be as small as 1 mm s(−)(1). Background velocities did not exceed 3 mm s(−)(1) as long as no moving animal was present in the measuring plane. The possible advantage for piscivorous predators of being able to detect and analyze fish-generated wakes is discussed.
APA, Harvard, Vancouver, ISO, and other styles
45

Schenck, T., and J. Jovanovic´. "Measurement of the Instantaneous Velocity Gradients in Plane and Axisymmetric Turbulent Wake Flows." Journal of Fluids Engineering 124, no. 1 (September 13, 2001): 143–53. http://dx.doi.org/10.1115/1.1428330.

Full text
Abstract:
All first-order spatial derivatives of the turbulent velocity fluctuations were measured using a pair of X hot-wire probes. Measurements were performed in the self-preserving region of a turbulent plane wake downstream of a cylinder and in an axisymmetric wake behind the sphere. Good spatial resolution of the measurements was ensured by choosing small values for the cylinder/sphere diameter and a low flow speed. Errors due to the finite hot-wire length and the wire and probe separation were analyzed using Wyngaard’s correction method. The derived corrections were verified experimentally. The measuring technique and the experimental results were systematically checked and compared with the results available in the literature. The assumptions of local isotropy and local axisymmetry were examined. Both investigated flows deviate only moderately from local isotropy and local axisymmetry. Support for the measured results is provided by plotting the data on an anisotropy invariant map. The budgets of the turbulent kinetic energy were computed from the measured data. In contrast to the results obtained in the plane wake, where the pressure transport is nearly negligible, in the axisymmetric wake it was found to play an important role and closely follows the estimate made by Lumley, uip¯/ρ≈−0.2q2ui¯.
APA, Harvard, Vancouver, ISO, and other styles
46

Hara, Tetsuya, Sigeru Miyoshi, and Petri Mähönen. "Formation Mechanism of Hierarchical Astronomical Objects in the Cosmic String Scheme." Symposium - International Astronomical Union 183 (1999): 249. http://dx.doi.org/10.1017/s0074180900132668.

Full text
Abstract:
Because of the angular deficit around cosmic string, cold dark matter(CDM) gets the velocity toward the plane of moving string and wake is formed there. It means that CDM body is cut by string, two blocks move each other, and the overlapped region becomes wake. The wake and/or block size is determined by the horizon size at zi when the wake is triggered by the string. The nonlinearity and/or thickness of the wake depends on the moving length of CDM toward the wake which depends on zi, the line density and velocity of string.
APA, Harvard, Vancouver, ISO, and other styles
47

Chernykh, G. G., A. G. Demenkov, O. V. Kaptsov, and A. V. Schmidt. "On Self-Similar Decay of a Plane Momentumless Turbulent Wake." Journal of Engineering Thermophysics 30, no. 4 (October 2021): 672–78. http://dx.doi.org/10.1134/s181023282104010x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Weygandt, James H., and Rabindra D. Mehta. "Effects of streamwise vorticity injection on a plane turbulent wake." AIAA Journal 33, no. 1 (January 1995): 86–93. http://dx.doi.org/10.2514/3.12336.

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

Marasli, B., F. H. Champagne, and I. J. Wygnanski. "Modal decomposition of velocity signals in a plane, turbulent wake." Journal of Fluid Mechanics 198, no. -1 (January 1989): 255. http://dx.doi.org/10.1017/s0022112089000121.

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

Antonia, R. A., and L. W. B. Browne. "Conventional and conditional Prandtl number in a turbulent plane wake." International Journal of Heat and Mass Transfer 30, no. 10 (October 1987): 2023–30. http://dx.doi.org/10.1016/0017-9310(87)90083-4.

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