Journal articles on the topic 'Vortex interactions'

To see the other types of publications on this topic, follow the link: Vortex interactions.

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 'Vortex interactions.'

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

Scott, R. K., and D. G. Dritschel. "Vortex–Vortex Interactions in the Winter Stratosphere." Journal of the Atmospheric Sciences 63, no. 2 (February 1, 2006): 726–40. http://dx.doi.org/10.1175/jas3632.1.

Full text
Abstract:
Abstract This paper examines the interaction of oppositely signed vortices in the compressible (non-Boussinesq) quasigeostrophic system, with a view to understanding vortex interactions in the polar winter stratosphere. A series of simplifying approximations leads to a two-vortex system whose dynamical properties are determined principally by two parameters: the ratio of the circulation of the vortices and the vertical separation of their centroids. For each point in this two-dimensional parameter space a family of equilibrium solutions exists, further parameterized by the horizontal separation of the vortex centroids, which are stable for horizontal separations greater than a critical value. The stable equilibria are characterized by vortex deformations that generally involve stronger deformations of the larger and/or lower of the two vortices. For smaller horizontal separations, the equilibria are unstable and a strongly nonlinear, time-dependent interaction takes place, typically involving the shedding of material from the larger vortex while the smaller vortex remains coherent. Qualitatively, the interactions resemble previous observations of certain stratospheric sudden warmings that involved the interaction of a growing anticyclonic circulation with the cyclonic polar vortex.
APA, Harvard, Vancouver, ISO, and other styles
2

ISHIKAWA, Hitoshi, Seiichiro IZAWA, Osamu MOCHIZUKI, and Masaru KIYA. "Vortex Ring-Vortex Tube Interactions." Transactions of the Japan Society of Mechanical Engineers Series B 68, no. 674 (2002): 2688–94. http://dx.doi.org/10.1299/kikaib.68.2688.

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

Peng, Di, and James W. Gregory. "Vortex dynamics during blade-vortex interactions." Physics of Fluids 27, no. 5 (May 2015): 053104. http://dx.doi.org/10.1063/1.4921449.

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

Rockwell, Donald. "VORTEX-BODY INTERACTIONS." Annual Review of Fluid Mechanics 30, no. 1 (January 1998): 199–229. http://dx.doi.org/10.1146/annurev.fluid.30.1.199.

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

Verzicco, R., and P. Orlandi. "Wall/Vortex-Ring Interactions." Applied Mechanics Reviews 49, no. 10 (October 1, 1996): 447–61. http://dx.doi.org/10.1115/1.3101985.

Full text
Abstract:
This review article presents a state-of-the-art review of the ring and wall interactions in the case of normal and oblique collisions. The different approaches used to study this flow and the results obtained are described and discussed. These techniques span from flow visualizations to LDV measurements, direct numerical simulations, particle-in-cell vortex methods and viscous and inviscid interactions. The relevance of these basic flows to the comprehension of wall-turbulence is also described. Finally, further developments, such as interaction with a grooved surface and with a deformable wall, and in Non-Newtonian fluids, are suggested.
APA, Harvard, Vancouver, ISO, and other styles
6

Doligalski, T. L., C. R. Smith, and J. D. A. Walker. "Vortex Interactions with Walls." Annual Review of Fluid Mechanics 26, no. 1 (January 1994): 573–616. http://dx.doi.org/10.1146/annurev.fl.26.010194.003041.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Kivshar, Yuri S., Alexander Nepomnyashchy, Vladimir Tikhonenko, Jason Christou, and Barry Luther-Davies. "Vortex-stripe soliton interactions." Optics Letters 25, no. 2 (January 15, 2000): 123. http://dx.doi.org/10.1364/ol.25.000123.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

FRITTS, DAVID C., STEVE ARENDT, and ØYVIND ANDREASSEN. "Vorticity dynamics in a breaking internal gravity wave. Part 2. Vortex interactions and transition to turbulence." Journal of Fluid Mechanics 367 (July 25, 1998): 47–65. http://dx.doi.org/10.1017/s0022112098001633.

Full text
Abstract:
A companion paper (Part 1) employed a three-dimensional numerical simulation to examine the vorticity dynamics of the initial instabilities of a breaking internal gravity wave in a stratified, sheared, compressible fluid. The present paper describes the vorticity dynamics that drive this flow to smaller-scale, increasingly isotropic motions at later times. Following the initial formation of discrete and intertwined vortex loops, the most important interactions are the self-interactions of single vortex tubes and the mutual interactions of multiple vortex tubes in close proximity. The initial formation of vortex tubes from the roll-up of localized vortex sheets gives the vortex tubes axial variations with both axisymmetric and azimuthal-wavenumber-2 components. The axisymmetric variations excite axisymmetric twist waves or Kelvin vortex waves which propagate along the tubes, drive axial flows, deplete the tubes' cores, and fragment the tubes. The azimuthal-wavenumber-2 variations excite m=2 twist waves on the vortex tubes, which undergo strong amplification and unravel single vortex tubes into pairs of intertwined helical tubes; the vortex tubes then burst or fragment. Reconnection often occurs among the remnants of such vortex fragmentation. A common mutual interaction is that of orthogonal vortex tubes, which causes mutual stretching and leads to long-lived structures. Such an interaction also sometimes creates an m=1 twist wave having an approximately steady helical form as well as a preferred sense of helicity. Interactions among parallel vortex tubes are less common, but include vortex pairing. Finally, the intensification and roll-up of weaker vortex sheets into new tubes occurs throughout the evolution. All of these vortex interactions result in a rapid cascade of energy and enstrophy toward smaller scales of motion.
APA, Harvard, Vancouver, ISO, and other styles
9

BAMBREY, ROSS R., JEAN N. REINAUD, and DAVID G. DRITSCHEL. "Strong interactions between two corotating quasi-geostrophic vortices." Journal of Fluid Mechanics 592 (November 14, 2007): 117–33. http://dx.doi.org/10.1017/s0022112007008373.

Full text
Abstract:
In this paper we investigate the interaction between two corotating quasi-geostrophic vortices. The initially ellipsoidal vortices are separated horizontally by a distance corresponding to the margin of stability, as determined from an ellipsoidal analysis. The subsequent interaction depends on four parameters: the vortex volume ratio, the vertical centroid separation, and the height-to-width aspect ratios of each vortex. The most commonly observed strong interaction is partial merger, where only part of the weaker vortex is incorporated into the stronger one or cast into filamentary debris. Despite the proliferation of small-scale filamentary structure during many vortex interactions, on average the self-induced vortex energy exhibits an ‘inverse cascade’ to larger scales, broadly consistent with spectral theories of turbulence. Curiously, we observe that a range of intermediate-scale vortices are preferentially sheared out during the interactions, leaving two main populations of large and small vortices.
APA, Harvard, Vancouver, ISO, and other styles
10

Roenby, Johan, and Hassan Aref. "Chaos in body–vortex interactions." Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 466, no. 2119 (February 24, 2010): 1871–91. http://dx.doi.org/10.1098/rspa.2009.0619.

Full text
Abstract:
The model of body–vortex interactions, where the fluid flow is planar, ideal and unbounded, and the vortex is a point vortex, is studied. The body may have a constant circulation around it. The governing equations for the general case of a freely moving body of arbitrary shape and mass density and an arbitrary number of point vortices are presented. The case of a body and a single vortex is then investigated numerically in detail. In this paper, the body is a homogeneous, elliptical cylinder. For large body–vortex separations, the system behaves much like a vortex pair regardless of body shape. The case of a circle is integrable. As the body is made slightly elliptic, a chaotic region grows from an unstable relative equilibrium of the circle-vortex case. The case of a cylindrical body of any shape moving in fluid otherwise at rest is also integrable. A second transition to chaos arises from the limit between rocking and tumbling motion of the body known in this case. In both instances, the chaos may be detected both in the body motion and in the vortex motion. The effect of increasing body mass at a fixed body shape is to damp the chaos.
APA, Harvard, Vancouver, ISO, and other styles
11

TANG, S. K., and N. W. M. KO. "Sound sources in the interactions of two inviscid two-dimensional vortex pairs." Journal of Fluid Mechanics 419 (September 25, 2000): 177–201. http://dx.doi.org/10.1017/s0022112000001294.

Full text
Abstract:
The sources of sound during the interactions of two identical two-dimensional inviscid vortex pairs are investigated numerically by using the vortex sound theory and the method of contour dynamics. The sound sources are identified and then separated into two independent components, which represent the contributions from the vortex centroid dynamics and the microscopic vortex core dynamics. Results show that the sound generation mechanism of the latter is independent of the type of vortex pair interaction, while that of the former depends on the jerks, accelerations and vortex forces on the vortex pairs. The power developed by the vortex forces is found to be important in the generation of sound when the vortex cores are severely deformed and their centroids are close to each other. The isolated source terms also explain the appearance of wavy oscillations on the time variations of the sound source strengths in the vortex ring and the two-dimensional vortex interaction systems.
APA, Harvard, Vancouver, ISO, and other styles
12

Chernodub, M. "Vortex profiles and vortex interactions at the electroweak crossover." Nuclear Physics B - Proceedings Supplements 83-84, no. 1-3 (March 2000): 571–73. http://dx.doi.org/10.1016/s0920-5632(00)00318-2.

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

Chernodub, M. N., E. M. Ilgenfritz, and A. Schiller. "Vortex profiles and vortex interactions at the electroweak crossover." Nuclear Physics B - Proceedings Supplements 83-84 (April 2000): 571–73. http://dx.doi.org/10.1016/s0920-5632(00)91741-9.

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

Renard, P. H., D. Thévenin, J. C. Rolon, and S. Candel. "Dynamics of flame/vortex interactions." Progress in Energy and Combustion Science 26, no. 3 (June 2000): 225–82. http://dx.doi.org/10.1016/s0360-1285(00)00002-2.

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

Rutland, Christopher J., and Joel H. Ferziger. "Simulations of flame-vortex interactions." Combustion and Flame 84, no. 3-4 (April 1991): 343–60. http://dx.doi.org/10.1016/0010-2180(91)90011-y.

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

Cutler, A. D., and P. Bradshaw. "Strong vortex/boundary layer interactions." Experiments in Fluids 14, no. 5 (April 1993): 321–32. http://dx.doi.org/10.1007/bf00189490.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Cutler, A. D., and P. Bradshaw. "Strong vortex/boundary layer interactions." Experiments in Fluids 14, no. 6 (May 1993): 393–401. http://dx.doi.org/10.1007/bf00190193.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Chorin, Alexandre Joel. "Hairpin removal in vortex interactions." Journal of Computational Physics 91, no. 1 (November 1990): 1–21. http://dx.doi.org/10.1016/0021-9991(90)90001-h.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Chorin, Alexander J. "Hairpin removal in vortex interactions." Journal of Computational Physics 87, no. 2 (April 1990): 496. http://dx.doi.org/10.1016/0021-9991(90)90272-3.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Dritschel, David G. "A general theory for two-dimensional vortex interactions." Journal of Fluid Mechanics 293 (June 25, 1995): 269–303. http://dx.doi.org/10.1017/s0022112095001716.

Full text
Abstract:
A general theory for two-dimensional vortex interactions is developed from the observation that, under slowly changing external influences, an individual vortex evolves through a series of equilibrium states until such a state proves unstable. Once an unstable equilibrium state is reached, a relatively fast unsteady evolution ensues, typically involving another nearby vortex. During this fast unsteady evolution, a fraction of the original coherent circulation is lost to filamentary debris, and, remarkably, the flow reorganizes into a set of quasi-steady stable vortices.The simplifying feature of the proposed theory is its use of adiabatic steadiness and marginal stability to determine the shapes and separation distance of vortices on the brink of an inelastic interaction. As a result, the parameter space for the inelastic interaction of nearby vortices is greatly reduced. In the case of two vortex patches, which is the focus of the present work, inelastic interactions depend only on a single parameter: the area ratio of the two vortices (taking the vorticity magnitude inside each to be equal). Without invoking adiabatic steadiness and marginal stability, one would have to contend with the additional parameters of vortex separation and shape, and the latter is actually an infinitude of parameters.
APA, Harvard, Vancouver, ISO, and other styles
21

KUO, ALLEN C., and LORENZO M. POLVANI. "Wave–vortex interaction in rotating shallow water. Part 1. One space dimension." Journal of Fluid Mechanics 394 (September 10, 1999): 1–27. http://dx.doi.org/10.1017/s0022112099005534.

Full text
Abstract:
Using a physical space (i.e. non-modal) approach, we investigate interactions between fast inertio-gravity (IG) waves and slow balanced flows in a shallow rotating fluid. Specifically, we consider a train of IG waves impinging on a steady, exactly balanced vortex. For simplicity, the one-dimensional problem is studied first; the limitations of one-dimensionality are offset by the ability to define balance in an exact way. An asymptotic analysis of the problem in the small-amplitude limit is performed to demonstrate the existence of interactions. It is shown that these interactions are not confined to the modification of the wave field by the vortex but, more importantly, that the waves are able to alter in a non-trivial way the potential vorticity associated with that vortex. Interestingly, in this one-dimensional problem, once the waves have traversed the vortex region and have propagated away, the vortex exactly recovers its initial shape and thus bears no signature of the interaction. Furthermore, we prove this last result in the case of arbitrary vortex and wave amplitudes. Numerical integrations of the full one-dimensional shallow-water equations in strongly nonlinear regimes are also performed: they confirm that time-dependent interactions exist and increase with wave amplitude, while at the final state the vortex bears no sign of the interaction. In addition, they reveal that cyclonic vortices interact more strongly with the wave field than anticyclonic ones.
APA, Harvard, Vancouver, ISO, and other styles
22

HULSHOFF, S. J., A. HIRSCHBERG, and G. C. J. HOFMANS. "Sound production of vortex–nozzle interactions." Journal of Fluid Mechanics 439 (July 23, 2001): 335–52. http://dx.doi.org/10.1017/s0022112001004554.

Full text
Abstract:
The factors which affect the sound production of a vortex as it passes through a nozzle are investigated at both low and high Mach numbers using time-accurate inviscid-flow computations. Vortex circulation, initial position, and mean-flow Mach number are shown to be the primary factors which influence the amplitude and phase of the sound produced. Nozzle geometry and distribution of vorticity are also shown to play significant roles in determining the detailed form of the signal. Additionally, it is shown that solution bifurcations are possible at sufficiently large values of vortex circulation. Comparisons are made between sound signals computed directly using a numerical method for the Euler equations and predictions obtained using a compressible vortex-sound analogy coupled with a compact-source assumption for the computation of vorticity dynamics. The results confirm that the latter approach is accurate for a range of problems with low mean-flow Mach numbers. At higher Mach numbers, however, the non-compactness of the source becomes apparent, resulting in significant changes to the character of the signal which cannot be predicted using the analogy-based approach. Implications for the construction of simplified models of vortex sound in solid-rocket nozzles are discussed.
APA, Harvard, Vancouver, ISO, and other styles
23

Kuo, Hung-Chi, Wayne H. Schubert, Chia-Ling Tsai, and Yu-Fen Kuo. "Vortex Interactions and Barotropic Aspects of Concentric Eyewall Formation." Monthly Weather Review 136, no. 12 (December 1, 2008): 5183–98. http://dx.doi.org/10.1175/2008mwr2378.1.

Full text
Abstract:
Abstract Concentric eyewall formation can be idealized as the interaction of a tropical cyclone core with nearby weaker vorticity of various spatial scales. This paper considers barotropic aspects of concentric eyewall formation from modified Rankine vortices. In this framework, the following parameters are found to be important in concentric eyewall formation: vorticity strength ratio, separation distance, companion vortex size, and core vortex skirt parameter. A vorticity skirt on the core vortex affects the filamentation dynamics in two important ways. First, the vorticity skirt lengthens the filamentation time, and therefore slows moat formation in the region just outside the radius of maximum wind. Second, at large radii, a skirted core vortex induces higher strain rates than a corresponding Rankine vortex and is thus more capable of straining out the vorticity field far from the core. Calculations suggest that concentric structures result from binary interactions when the small vortex is at least 4–6 times as strong as the larger companion vortex. An additional requirement is that the separation distance between the edges of the two vortices be less than 6–7 times the smaller vortex radius. Broad moats form when the initial companion vortex is small, the vorticity skirt outside the radius of maximum wind is small, and the strength ratio is large. In concentric cases, an outer vorticity ring develops when the initial companion vortex is large, the vorticity skirt outside the radius of maximum wind is small, and the strength ratio is not too large. In general, when the companion vortex is 3 times as strong as the core vortex and the separation distance is 4–6 times the radius of the smaller vortex, a core vortex with a vorticity skirt produces concentric structures. In contrast, a Rankine vortex produces elastic interaction in this region of parameter space. Thus, a Rankine vortex of sufficient strength favors the formation of a concentric structure closer to the core vortex, while a skirted vortex of sufficient strength favors the formation of concentric structures farther from the core vortex. This may explain satellite microwave observations that suggest a wide range of radii for concentric eyewalls.
APA, Harvard, Vancouver, ISO, and other styles
24

Nair, Aditya G., and Kunihiko Taira. "Network-theoretic approach to sparsified discrete vortex dynamics." Journal of Fluid Mechanics 768 (March 10, 2015): 549–71. http://dx.doi.org/10.1017/jfm.2015.97.

Full text
Abstract:
We examine discrete vortex dynamics in two-dimensional flow through a network-theoretic approach. The interaction of the vortices is represented with a graph, which allows the use of network-theoretic approaches to identify key vortex-to-vortex interactions. We employ sparsification techniques on these graph representations based on spectral theory to construct sparsified models and evaluate the dynamics of vortices in the sparsified set-up. Identification of vortex structures based on graph sparsification and sparse vortex dynamics is illustrated through an example of point-vortex clusters interacting amongst themselves. We also evaluate the performance of sparsification with increasing number of point vortices. The sparsified-dynamics model developed with spectral graph theory requires a reduced number of vortex-to-vortex interactions but agrees well with the full nonlinear dynamics. Furthermore, the sparsified model derived from the sparse graphs conserves the invariants of discrete vortex dynamics. We highlight the similarities and differences between the present sparsified-dynamics model and reduced-order models.
APA, Harvard, Vancouver, ISO, and other styles
25

KALKHORAN, I. M., M. K. SMART, and F. Y. WANG. "Supersonic vortex breakdown during vortex/cylinder interaction." Journal of Fluid Mechanics 369 (August 25, 1998): 351–80. http://dx.doi.org/10.1017/s0022112098001566.

Full text
Abstract:
The head-one interaction of a supersonic streamwise vortex with a circular cylinder reveals a vortex breakdown similar in many ways to that of incompressible vortex breakdown. In particular, the dramatic flow reorganization observed during the interaction resembles the conical vortex breakdown reported by Sarpkaya (1995) at high Reynolds number. In the present study, vortex breakdown is brought about when moderate and strong streamwise vortices encounter the bow shock in front of a circular cylinder at Mach 2.49. The main features of the vortex/cylinder interaction are the formation of a blunt-nosed conical shock with apex far upstream of the undisturbed shock stand-off distance, and a vortex core which responds to passage through the apex of the conical shock by expanding into a turbulent conical flow structure. The geometry of the expanding vortex core as well as the location of the conical shock apex are seen to be strong functions of the incoming vortex strength and the cylinder diameter. A salient feature of the supersonic vortex breakdown is the formation of an entropy-shear layer, which separates an interior subsonic zone containing the burst vortex from the surrounding supersonic flow. In keeping with the well-established characteristics of the low-speed vortex breakdown, a region of reversed flow is observed inside the turbulent subsonic zone. The steady vortex/cylinder interaction flow fields generated in the current study exhibit many characteristics of the unsteady vortex distortion patterns previously observed during normal shock wave/vortex interactions. This similarity of the instantaneous flow structure indicates that the phenomenon previously called vortex distortion by Kalkhoran et al. (1996) is a form of supersonic vortex breakdown.
APA, Harvard, Vancouver, ISO, and other styles
26

Cenedese, Claudia, Robert E. Todd, Glen G. Gawarkiewicz, W. Brechner Owens, and Andrey Y. Shcherbina. "Offshore Transport of Shelf Waters through Interaction of Vortices with a Shelfbreak Current." Journal of Physical Oceanography 43, no. 5 (May 1, 2013): 905–19. http://dx.doi.org/10.1175/jpo-d-12-0150.1.

Full text
Abstract:
Abstract Interactions between vortices and a shelfbreak current are investigated, with particular attention to the exchange of waters between the continental shelf and slope. The nonlinear, three-dimensional interaction between an anticyclonic vortex and the shelfbreak current is studied in the laboratory while varying the ratio ε of the maximum azimuthal velocity in the vortex to the maximum alongshelf velocity in the shelfbreak current. Strong interactions between the shelfbreak current and the vortex are observed when ε > 1; weak interactions are found when ε < 1. When the anticyclonic vortex comes in contact with the shelfbreak front during a strong interaction, a streamer of shelf water is drawn offshore and wraps anticyclonically around the vortex. Measurements of the offshore transport and identification of the particle trajectories in the shelfbreak current drawn offshore from the vortex allow quantification of the fraction of the shelfbreak current that is deflected onto the slope; this fraction increases for increasing values of ε. Experimental results in the laboratory are strikingly similar to results obtained from observations in the Middle Atlantic Bight (MAB); after proper scaling, measurements of offshore transport and offshore displacement of shelf water for vortices in the MAB that span a range of values of ε agree well with laboratory predictions.
APA, Harvard, Vancouver, ISO, and other styles
27

Forster, Kyle J., Sammy Diasinos, Graham Doig, and Tracie J. Barber. "Large eddy simulation of transient upstream/downstream vortex interactions." Journal of Fluid Mechanics 862 (January 9, 2019): 227–60. http://dx.doi.org/10.1017/jfm.2018.949.

Full text
Abstract:
Experimentally validated large eddy simulations were performed on two NACA0012 vanes at various lateral offsets to observe the transient effects of the near field interactions between two streamwise vortices. The vanes were separated in the streamwise direction, allowing the upstream vortex to impact on the downstream geometry. These vanes were evaluated at an angle of incidence of $8^{\circ }$ and a Reynolds number of 70 000, with rear vane angle reversed to create a co-rotating or counter-rotating vortex pair. The downstream vortex merged with the upstream in the co-rotating condition, driven by the suppression of one of the tip vortices of the downstream vane. At close proximity to the pressure side, the vane elongated the upstream vortex, resulting in it being the weakened and merging into the downstream vortex. This produced a transient production of bifurcated vortices in the wake region. The downstream vortex of the co-rotating pair experienced faster meandering growth, with position oscillations equalising between the vortices. The position oscillation was determined to be responsible for statistical variance in the merging location, with variation in vortex separation causing the vortices at a single plane to merge and separate in a time-dependent manner. In the counter-rotating condition position oscillations were found to be larger, with higher growth, but less uniform periodicity. It was found that the circulation transfer between the vortices was linked to the magnitude of their separation, with high separation fluctuations weakening the upstream vortex and strengthening the downstream vortex. In the case of upstream vortex impingement on the downstream vane, the upstream vortex was found to bifurcate, with a four vortex system being formed by interactions with the shear layer. This eventually resulted in a single dominant vortex, which did not magnify its oscillation amplitudes as it travelled downstream due to the destruction of the interacting vortices.
APA, Harvard, Vancouver, ISO, and other styles
28

Folz, Patrick J. R., and Keiko K. Nomura. "A quantitative assessment of viscous asymmetric vortex pair interactions." Journal of Fluid Mechanics 829 (September 14, 2017): 1–30. http://dx.doi.org/10.1017/jfm.2017.527.

Full text
Abstract:
The interactions of two like-signed vortices in viscous fluid are investigated using two-dimensional numerical simulations performed across a range of vortex strength ratios, $\unicode[STIX]{x1D6EC}=\unicode[STIX]{x1D6E4}_{1}/\unicode[STIX]{x1D6E4}_{2}\leqslant 1$, corresponding to vortices of circulation, $\unicode[STIX]{x1D6E4}_{i}$, with differing initial size and/or peak vorticity. In all cases, the vortices evolve by viscous diffusion before undergoing a primary convective interaction, which ultimately results in a single vortex. The post-interaction vortex is quantitatively evaluated in terms of an enhancement factor, $\unicode[STIX]{x1D700}=\unicode[STIX]{x1D6E4}_{end}/\unicode[STIX]{x1D6E4}_{2,start}$, which compares its circulation, $\unicode[STIX]{x1D6E4}_{end}$, to that of the stronger starting vortex, $\unicode[STIX]{x1D6E4}_{2,start}$. Results are effectively characterized by a mutuality parameter, $MP\equiv (S/\unicode[STIX]{x1D714})_{1}/(S/\unicode[STIX]{x1D714})_{2}$, where the ratio of induced strain rate, $S$, to peak vorticity, $\unicode[STIX]{x1D714}$, for each vortex, $(S/\unicode[STIX]{x1D714})_{i}$, is found to have a critical value, $(S/\unicode[STIX]{x1D714})_{cr}\approx 0.135$, above which core detrainment occurs. If $MP$ is sufficiently close to unity, both vortices detrain and a two-way mutual entrainment process leads to $\unicode[STIX]{x1D700}>1$, i.e. merger. In asymmetric interactions and mergers, generally one vortex dominates; the weak/no/strong vortex winner regimes correspond to $MP<,=,>1$, respectively. As $MP$ deviates from unity, $\unicode[STIX]{x1D700}$ decreases until a critical value, $MP_{cr}$ is reached, beyond which there is only a one-way interaction; one vortex detrains and is destroyed by the other, which dominates and survives. There is no entrainment and $\unicode[STIX]{x1D700}\sim 1$, i.e. only a straining out occurs. Although $(S/\unicode[STIX]{x1D714})_{cr}$ appears to be independent of Reynolds number, $MP_{cr}$ shows a dependence. Comparisons are made with available experimental data from Meunier (2001, PhD thesis, Université de Provence-Aix-Marseille I).
APA, Harvard, Vancouver, ISO, and other styles
29

Sakajo, Takashi, and Yuuki Shimizu. "Point vortex interactions on a toroidal surface." Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 472, no. 2191 (July 2016): 20160271. http://dx.doi.org/10.1098/rspa.2016.0271.

Full text
Abstract:
Owing to non-constant curvature and a handle structure, it is not easy to imagine intuitively how flows with vortex structures evolve on a toroidal surface compared with those in a plane, on a sphere and a flat torus. In order to cultivate an insight into vortex interactions on this manifold, we derive the evolution equation for N -point vortices from Green's function associated with the Laplace–Beltrami operator there, and we then formulate it as a Hamiltonian dynamical system with the help of the symplectic geometry and the uniformization theorem. Based on this Hamiltonian formulation, we show that the 2-vortex problem is integrable. We also investigate the point vortex equilibria and the motion of two-point vortices with the strengths of the same magnitude as one of the fundamental vortex interactions. As a result, we find some characteristic interactions between point vortices on the torus. In particular, two identical point vortices can be locally repulsive under a certain circumstance.
APA, Harvard, Vancouver, ISO, and other styles
30

Garmann, D. J., and M. R. Visbal. "Interactions of a streamwise-oriented vortex with a finite wing." Journal of Fluid Mechanics 767 (February 24, 2015): 782–810. http://dx.doi.org/10.1017/jfm.2015.51.

Full text
Abstract:
AbstractA canonical study is developed to investigate the unsteady interactions of a streamwise-oriented vortex impinging upon a finite surface using high-fidelity simulation. As a model problem, an analytically defined vortex superimposed on a free stream is convected towards an aspect-ratio-six ($\mathit{AR}=6$) plate oriented at an angle of ${\it\alpha}=4^{\circ }$ and Reynolds number of $\mathit{Re}=20\,000$ in order to characterize the unsteady modes of interaction resulting from different spanwise positions of the incoming vortex. Outboard, tip-aligned and inboard positioning are shown to produce three distinct flow regimes: when the vortex is positioned outboard of, but in close proximity to, the wingtip, it pairs with the tip vortex to form a dipole that propels itself away from the plate through mutual induction, and also leads to an enhancement of the tip vortex. When the incoming vortex is aligned with the wingtip, the tip vortex is initially strengthened by the proximity of the incident vortex, but both structures attenuate into the wake as instabilities arise in the pair’s feeding sheets from the entrainment of opposite-signed vorticity into either structure. Finally, when the incident vortex is positioned inboard of the wingtip, the vortex bifurcates in the time-mean sense with portions convecting above and below the wing, and the tip vortex is mostly suppressed. The time-mean bifurcation is actually a result of an unsteady spiralling instability in the vortex core that reorients the vortex as it impacts the leading edge, pinches off, and alternately attaches to either side of the wing. The increased effective angle of attack inboard of impingement enhances the three-dimensional recirculation region created by the separated boundary layer off the leading edge which draws fluid from the incident vortex inboard and diminishes its impact on the outboard section of the wing. The slight but remaining downwash present outboard of impingement reduces the effective angle of attack in that region, resulting in a small separation bubble on either side of the wing in the time-mean solution, effectively unloading the tip outboard of impingement and suppressing the tip vortex. All incident vortex positions provide substantial increases in the wing’s lift-to-drag ratio; however, significant sustained rolling moments also result. As the vortex is brought inboard, the rolling moment diminishes and eventually switches sign as the reduced outboard loading balances the augmented sectional lift inboard of impingement.
APA, Harvard, Vancouver, ISO, and other styles
31

Gemmell, Brad J., Kevin T. Du Clos, Sean P. Colin, Kelly R. Sutherland, and John H. Costello. "The most efficient metazoan swimmer creates a ‘virtual wall’ to enhance performance." Proceedings of the Royal Society B: Biological Sciences 288, no. 1942 (January 6, 2021): 20202494. http://dx.doi.org/10.1098/rspb.2020.2494.

Full text
Abstract:
It has been well documented that animals (and machines) swimming or flying near a solid boundary get a boost in performance. This ground effect is often modelled as an interaction between a mirrored pair of vortices represented by a true vortex and an opposite sign ‘virtual vortex’ on the other side of the wall. However, most animals do not swim near solid surfaces and thus near body vortex–vortex interactions in open-water swimmers have been poorly investigated. In this study, we examine the most energetically efficient metazoan swimmer known to date, the jellyfish Aurelia aurita , to elucidate the role that vortex interactions can play in animals that swim away from solid boundaries. We used high-speed video tracking, laser-based digital particle image velocimetry (dPIV) and an algorithm for extracting pressure fields from flow velocity vectors to quantify swimming performance and the effect of near body vortex–vortex interactions. Here, we show that a vortex ring (stopping vortex), created underneath the animal during the previous swim cycle, is critical for increasing propulsive performance. This well-positioned stopping vortex acts in the same way as a virtual vortex during wall-effect performance enhancement, by helping converge fluid at the underside of the propulsive surface and generating significantly higher pressures which result in greater thrust. These findings advocate that jellyfish can generate a wall-effect boost in open water by creating what amounts to a ‘virtual wall’ between two real, opposite sign vortex rings. This explains the significant propulsive advantage jellyfish possess over other metazoans and represents important implications for bio-engineered propulsion systems.
APA, Harvard, Vancouver, ISO, and other styles
32

Zhang, Yu, Joseph Pedlosky, and Glenn R. Flierl. "Shelf Circulation and Cross-Shelf Transport out of a Bay Driven by Eddies from an Open-Ocean Current. Part I: Interaction between a Barotropic Vortex and a Steplike Topography." Journal of Physical Oceanography 41, no. 5 (May 1, 2011): 889–910. http://dx.doi.org/10.1175/2010jpo4496.1.

Full text
Abstract:
Abstract This paper examines interaction between a barotropic point vortex and a steplike topography with a bay-shaped shelf. The interaction is governed by two mechanisms: propagation of topographic Rossby waves and advection by the forcing vortex. Topographic waves are supported by the potential vorticity (PV) jump across the topography and propagate along the step only in one direction, having higher PV on the right. Near one side boundary of the bay, which is in the wave propagation direction and has a narrow shelf, waves are blocked by the boundary, inducing strong out-of-bay transport in the form of detached crests. The wave–boundary interaction as well as out-of-bay transport is strengthened as the minimum shelf width is decreased. The two control mechanisms are related differently in anticyclone- and cyclone-induced interactions. In anticyclone-induced interactions, the PV front deformations are moved in opposite directions by the point vortex and topographic waves; a topographic cyclone forms out of the balance between the two opposing mechanisms and is advected by the forcing vortex into the deep ocean. In cyclone-induced interactions, the PV front deformations are moved in the same direction by the two mechanisms; a topographic cyclone forms out of the wave–boundary interaction but is confined to the coast. Therefore, anticyclonic vortices are more capable of driving water off the topography. The anticyclone-induced transport is enhanced for smaller vortex–step distance or smaller topography when the vortex advection is relatively strong compared to the wave propagation mechanism.
APA, Harvard, Vancouver, ISO, and other styles
33

George, A. R., and A. S. Lyrintzis. "Acoustics of transonic blade-vortex interactions." AIAA Journal 26, no. 7 (July 1988): 769–76. http://dx.doi.org/10.2514/3.9968.

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

Katz, Joseph. "Wing/vortex interactions and wing rock." Progress in Aerospace Sciences 35, no. 7 (October 1999): 727–50. http://dx.doi.org/10.1016/s0376-0421(99)00004-4.

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

Jaworski, J. W. "Sound from aeroelastic vortex–fibre interactions." Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 377, no. 2159 (October 14, 2019): 20190071. http://dx.doi.org/10.1098/rsta.2019.0071.

Full text
Abstract:
The motion of a line vortex moving past a one-dimensional flexible fibre is examined theoretically. A Schwarz–Christoffel conformal mapping enables the analytical solution of the potential flow field and its hydrodynamic moment on the flexible fibre, which is composed of a rigid segment constrained to angular motions on a wedge. The hydroelastic coupling of the vortex path and fibre motion affects the noise signature, which is evaluated for the special case of acoustically compact fibres embedded in a half plane. Results from this analysis attempt to address how the coupled interactions between vortical sources and flexible barbules on the upper surface of owl wings may contribute to their acoustic stealth. The analytical formulation is also amenable to application to vortex sound prediction from flexible trailing edges provided that an appropriate acoustic Green's function can be determined. This article is part of the theme issue ‘Frontiers of aeroacoustics research: theory, computation and experiment’.
APA, Harvard, Vancouver, ISO, and other styles
36

INOUE, OSAMU, and YUJI HATTORI. "Sound generation by shock–vortex interactions." Journal of Fluid Mechanics 380 (February 10, 1999): 81–116. http://dx.doi.org/10.1017/s0022112098003565.

Full text
Abstract:
Two-dimensional, unsteady, compressible flow fields produced by the interactions between a single vortex or a pair of vortices and a shock wave are simulated numerically. The Navier–Stokes equations are solved by a finite difference method. The sixth-order-accurate compact Padé scheme is used for spatial derivatives, together with the fourth-order-accurate Runge–Kutta scheme for time integration. The detailed mechanics of the flow fields at an early stage of the interactions and the basic nature of the near-field sound generated by the interactions are studied. The results for both a single vortex and a pair of vortices suggest that the generation and the nature of sounds are closely related to the generation of reflected shock waves. The flow field differs significantly when the pair of vortices moves in the same direction as the shock wave than when opposite to it.
APA, Harvard, Vancouver, ISO, and other styles
37

Wakelin, S. "Vortex ring interactions II. Inviscid models." Quarterly Journal of Mechanics and Applied Mathematics 49, no. 2 (May 1, 1996): 287–309. http://dx.doi.org/10.1093/qjmam/49.2.287.

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

Xue, Y., and A. S. Lyrintzis. "Transonic Blade-Vortex Interactions: Noise Reduction." Journal of Aircraft 30, no. 3 (May 1993): 408–11. http://dx.doi.org/10.2514/3.56888.

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

Rogberg, Peter, and David G. Dritschel. "Mixing in two-dimensional vortex interactions." Physics of Fluids 12, no. 12 (December 2000): 3285–88. http://dx.doi.org/10.1063/1.1320838.

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

HaoJie, YAN, QIN FengHua, and LUO XiSheng. "Numerical study on vortex-nozzle interactions." SCIENTIA SINICA Physica, Mechanica & Astronomica 48, no. 12 (October 8, 2018): 124701. http://dx.doi.org/10.1360/sspma2018-00085.

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

Chorin, Alexandre Joel. "Hairpin Removal in Vortex Interactions II." Journal of Computational Physics 107, no. 1 (July 1993): 1–9. http://dx.doi.org/10.1006/jcph.1993.1120.

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

Koumoutsakos, Petros. "Active control of vortex–wall interactions." Physics of Fluids 9, no. 12 (December 1997): 3808–16. http://dx.doi.org/10.1063/1.869515.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Zhang, M. M., L. Cheng, and Y. Zhou. "Closed-loop controlled vortex-airfoil interactions." Physics of Fluids 18, no. 4 (April 2006): 046102. http://dx.doi.org/10.1063/1.2189287.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Ferrell, Richard A., and Bumsoo Kyung. "Roton-vortex interactions in superfluid helium." Physical Review Letters 67, no. 8 (August 19, 1991): 1003–6. http://dx.doi.org/10.1103/physrevlett.67.1003.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Leonard, A., and K. Chua. "Three-dimensional interactions of vortex tubes." Physica D: Nonlinear Phenomena 37, no. 1-3 (July 1989): 490–96. http://dx.doi.org/10.1016/0167-2789(89)90153-x.

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

Grasso, F., and S. Pirozzoli. "Shock-Wave-Vortex Interactions: Shock and Vortex Deformations, and Sound Production." Theoretical and Computational Fluid Dynamics 13, no. 6 (March 1, 2000): 421–56. http://dx.doi.org/10.1007/s001620050121.

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

Nguyen, Van Luc, Tomohiro Degawa, and Tomomi Uchiyama. "Numerical simulation of the interaction between a vortex ring and a bubble plume." International Journal of Numerical Methods for Heat & Fluid Flow 29, no. 9 (September 2, 2019): 3192–224. http://dx.doi.org/10.1108/hff-12-2018-0734.

Full text
Abstract:
Purpose This paper aims to provide discussions of a numerical method for bubbly flows and the interaction between a vortex ring and a bubble plume. Design/methodology/approach Small bubbles are released into quiescent water from a cylinder tip. They rise under the buoyant force, forming a plume. A vortex ring is launched vertically upward into the bubble plume. The interactions between the vortex ring and the bubble plume are numerically simulated using a semi-Lagrangian–Lagrangian approach composed of a vortex-in-cell method for the fluid phase and a Lagrangian description of the gas phase. Findings A vortex ring can transport the bubbles surrounding it over a distance significantly depending on the correlative initial position between the bubbles and the core center. The motion of some bubbles is nearly periodic and gradually extinguishes with time. These bubble trajectories are similar to two-dimensional-helix shapes. The vortex is fragmented into multiple regions with high values of Q, the second invariant of velocity gradient tensor, settling at these regional centers. The entrained bubbles excite a growth rate of the vortex ring's azimuthal instability with a formation of the second- and third-harmonic oscillations of modes of 16 and 24, respectively. Originality/value A semi-Lagrangian–Lagrangian approach is applied to simulate the interactions between a vortex ring and a bubble plume. The simulations provide the detail features of the interactions.
APA, Harvard, Vancouver, ISO, and other styles
48

Mishra, Prasad, Renganathan Sudharshan, and Kumar Ezhil. "Numerical study of flame/vortex interactions in 2-D Trapped Vortex Combustor." Thermal Science 18, no. 4 (2014): 1373–87. http://dx.doi.org/10.2298/tsci111006162m.

Full text
Abstract:
The interactions between flame and vortex in a 2-D Trapped Vortex Combustor are investigated by simulating the Reynolds Averaged Navier Stokes (RANS) equations, for the following five cases namely (i) non-reacting (base) case, (ii) post-vortex ignition without premixing, (iii) post-vortex ignition with premixing, (iv) pre-vortex ignition without premixing and (v) pre-vortex ignition with premixing. For the post-vortex ignition without premixing case, the reactants are mixed well in the cavity resulting in a stable ?C? shaped flame along the vortex edge. Further, there is insignificant change in the vorticity due to chemical reactions. In contrast, for the pre-vortex ignition case (no premixing); the flame gets stabilized at the interface of two counter rotating vortices resulting in reduced reaction rates. There is a noticeable change in the location and size of the primary vortex as compared to case (ii). When the mainstream air is premixed with fuel, there is a further reduction in the reaction rates and thus structure of cavity flame gets altered significantly for case (v). Pilot flame established for cases (ii) and (iii) are well shielded from main flow and hence the flame structure and reaction rates do not change appreciably. Hence, it is expected that cases (ii) and (iii) can perform well over a wide range of operating conditions.
APA, Harvard, Vancouver, ISO, and other styles
49

Lentink, David, GertJan F. Van Heijst, Florian T. Muijres, and Johan L. Van Leeuwen. "Vortex interactions with flapping wings and fins can be unpredictable." Biology Letters 6, no. 3 (February 3, 2010): 394–97. http://dx.doi.org/10.1098/rsbl.2009.0806.

Full text
Abstract:
As they fly or swim, many animals generate a wake of vortices with their flapping fins and wings that reveals the dynamics of their locomotion. Previous studies have shown that the dynamic interaction of vortices in the wake with fins and wings can increase propulsive force. Here, we explore whether the dynamics of the vortex interactions could affect the predictability of propulsive forces. We studied the dynamics of the interactions between a symmetrically and periodically pitching and heaving foil and the vortices in its wake, in a soap-film tunnel. The phase-locked movie sequences reveal that abundant chaotic vortex-wake interactions occur at high Strouhal numbers. These high numbers are representative for the fins and wings of near-hovering animals. The chaotic wake limits the forecast horizon of the corresponding force and moment integrals. By contrast, we find periodic vortex wakes with an unlimited forecast horizon for the lower Strouhal numbers (0.2–0.4) at which many animals cruise. These findings suggest that swimming and flying animals could control the predictability of vortex-wake interactions, and the corresponding propulsive forces with their fins and wings.
APA, Harvard, Vancouver, ISO, and other styles
50

LE DIZÈS, STÉPHANE, and ALBERTO VERGA. "Viscous interactions of two co-rotating vortices before merging." Journal of Fluid Mechanics 467 (September 24, 2002): 389–410. http://dx.doi.org/10.1017/s0022112002001532.

Full text
Abstract:
The viscous evolution of two co-rotating vortices is analysed using direct two-dimensional numerical simulations of the Navier–Stokes equations. The article focuses on vortex interaction regimes before merging. Two parameters are varied: a steepness parameter n which measures the steepness of the initial vorticity profiles in a given family of profiles, and the Reynolds number Re (between 500 and 16 000). Two distinct relaxation processes are identified. The first one is non-viscous and corresponds to a rapid adaptation of each vortex to the external (strain) field generated by the other vortex. This adaptation process, which is profile dependent, is described and explained using the damped Kelvin modes of each vortex. The second relaxation process is a slow diffusion phenomenon. It is similar to the relaxation of any non-Gaussian axisymmetrical vortex towards the Gaussian. The quasi-stationary solution evolves on a viscous-time scale toward a single attractive solution which corresponds to the evolution from two initially Gaussian vortices. The attractive solution is analysed in detail up to the merging threshold a/b ≈ 0.22 where a and b are the vortex radius and the separation distance respectively. The vortex core deformations are quantified and compared to those induced by a single vortex in a rotating strain field. A good agreement with the asymptotic predictions is demonstrated for the eccentricity of vortex core streamlines. A weak anomalous Reynolds number dependence of the solution is also identified. This dependence is attributed to the advection–diffusion of vorticity towards the hyperbolic points of the system and across the separatrix connecting these points. A Re1/3 scaling for the vorticity at the central hyperbolic point is obtained. These findings are discussed in the context of a vortex merging criterion.
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