Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-wzw2p Total loading time: 0 Render date: 2024-05-17T09:10:44.556Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  20 February 2023

George Haller
Affiliation:
ETH Zurich
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Transport Barriers and Coherent Structures in Flow Data
Advective, Diffusive, Stochastic and Active Methods
, pp. 390 - 405
Publisher: Cambridge University Press
Print publication year: 2023

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Abernathey, R., and Haller, G. (2018). Transport by Lagrangian vortices in the Eastern Pacific. J. Phys. Oceanogr., 48, 667685.CrossRefGoogle Scholar
Abraham, R., Marsden, J. E., and Ratiu, T. (1988). Manifolds, Tensor Analysis, and Applications. Springer, New York, NY.Google Scholar
Adrian, R. J., Meinhart, C. D., and Tomkins, C. D. (2000). Vortex organization in the outer region of the turbulent boundary layer. J. Fluid Mech., 422, 154.CrossRefGoogle Scholar
Aksamit, N. O., and Haller, G. (2022). Objective momentum barriers in wall turbulence. J. Fluid Mech., 941, A3.Google Scholar
Aksamit, N. O., Sapsis, T., and Haller, G. (2020). Machine-learning mesoscale and submesoscale surface dynamics from Lagrangian ocean drifter trajectories. J. Phys. Oceanogr., 50(5), 11791196.Google Scholar
Aksamit, N. O., Kravitz, B., MacMartin, D. G., and Haller, G. (2022). Harnessing stratospheric diffusion barriers for enhanced climate geoengineering. Atmosph. Chem. and Phys., 21(11), 88458861.Google Scholar
Alam, M.-R., Liu, W., and Haller, G. (2006). Closed-loop separation control: An analytic approach. Phys. Fluids, 18(4), 043601.CrossRefGoogle Scholar
Alexander, P. (2004). High order computation of the history term in the equation of motion for a spherical particle in a fluid. J. Scientific Comp., 21, 129143.CrossRefGoogle Scholar
Andrade-Canto, F., Karrasch, D., and Beron-Vera, F. J. (2020). Genesis, evolution, and apocalypse of Loop Current rings. Phys. Fluids, 32(11), 116603.Google Scholar
Anghan, C., Dave, S., Saincher, S., and Banerjee, J. (2019). Direct numerical simulation of transitional and turbulent round jets: Evolution of vortical structures and turbulence budget. Phys. Fluids, 31, 053606.CrossRefGoogle Scholar
Angilella, J.-R. (2007). Asymptotic analysis of chaotic particle sedimentation and trapping in the vicinity of a vertical upward streamline. Phys. Fluids, 19, 073302.CrossRefGoogle Scholar
Antuono, M. (2020). Tri-periodic fully three-dimensional analytic solutions for the Navier–Stokes equations. J. Fluid Mech., 890, A23.Google Scholar
Aref, H. (1984). Stirring by chaotic advection. J. Fluid Mech., 143, 121.Google Scholar
Aref, H. (2002). The development of chaotic advection. Phys. Fluids, 14, 13151325.Google Scholar
Arnold, V. (1965). Sur la topologie des écoulements stationnaires des fluides parfaits. C. R. Acad. Sci. Paris, 261, 1720.Google Scholar
Arnold, V. (1966). Sur la géométrie différentielle des groupes de Lie de dimension infinie et ses applications à l’hydrodynamique des fluides parfaits. Annales de l’Institut Fourier, 16(1), 319361.Google Scholar
Arnold, V. I. (1978). Ordinary Differential Equations. MIT Press, Cambridge, MA.Google Scholar
Arnold, V. I. (1989). Mathematical Methods of Classical Mechanics. Springer-Verlag, New York, NY.CrossRefGoogle Scholar
Arnold, V. I., and Keshin, B. A. (1998). Topological Methods of Hydrodynamics. Springer-Verlag, New York, NY.Google Scholar
Artale, V., Boffetta, G., Celani, A., Cencini, M., and Vulpiani, A. (1997). Dispersion of passive tracers in closed basins: Beyond the diffusion coefficient. Phys. Fluids, 9(11), 31623171.Google Scholar
Astarita, G. (1979). Objective and generally applicable criteria for flow classification. J. Non-Newtonian Fluid Mech., 6, 6976.Google Scholar
Aulbach, B., and Wanner, T. (2000). The Hartman–Grobman theorem for Carathéodory-type differential equations in Banach spaces. Nonlin. Anal. Theory Methods Appl., 40(1), 91104.CrossRefGoogle Scholar
Aurell, E., Boffetta, G., Crisanti, A., Paladin, G., and Vulpiani, A. (1997). Predictability in the large: An extension of the concept of Lyapunov exponent. J. Phys. A, 30, 126.CrossRefGoogle Scholar
Babiano, A., Basdevant, C., Legras, B., and Sadourny, R. (1987). Vorticity and passive-scalar dynamics in two-dimensional turbulence. J. Fluid Mech., 183, 379397.Google Scholar
Babiano, A., Cartwright, J. H. E., Piro, O., and Provenzale, A. (2000). Dynamics of a small neutrally buoyant sphere in a fluid and targeting in Hamiltonian systems. Phys. Rev. Lett., 84, 57645767.Google Scholar
Balasuriya, S. (2016). Barriers and Transport in Unsteady Flows: A Melnikov Approach. SIAM, Philadelphia, PA.Google Scholar
Balasuriya, S., Ouellette, N. T., and Rypina, I. I. (2018). Generalized Lagrangian coherent structures. Physica D, 372, 3151.CrossRefGoogle Scholar
Bandle, C. (1980). Isoperimetric Inequalities and Applications. Pitman, Boston, MA., London.Google Scholar
Banisch, R., and Koltai, P. (2017). Understanding the geometry of transport: Diffusion maps for Lagrangian trajectory data unravel coherent sets. Chaos, 27(3), 035804.Google Scholar
Barbato, D., Berselli, L.-C., and Grisanti, C. R. (2007). Analytical and numerical results for the rational large eddy simulation model. J. Math. Fluid Mech., 9, 4474.Google Scholar
Basar, Y., and Weichert, D. (2000). Nonlinear Continuum Mechanics of Solids: Fundamental Mathematical and Physical Concepts. Springer-Verlag, Berlin, Germany.CrossRefGoogle Scholar
Basdevant, C., and Philopovitch, T. (1994). On the validity of the “Weiss criterion” in two-dimensional turbulence. Physica D, 73, 1730.Google Scholar
Beal, L., De Ruijter, W., Biastoch, A., et al. (2011). On the role of the Agulhas system in ocean circulation and climate. Nature, 472, 429436.CrossRefGoogle ScholarPubMed
Beem, J. K., Ehrlich, P. E., and Easley, K. L. (1996). Global Lorentzian Geometry. Taylor & Francis, New York, NY.Google Scholar
Benczik, I. J., Toroczkai, Z., and Tél, T. (2002). Selective sensitivity of open chaotic flows on inertial tracer advection: Catching particles with a stick. Phys. Rev. Lett., 89, 164501.CrossRefGoogle ScholarPubMed
Bennett, A. (2006). Lagrangian Fluid Dynamics. Cambridge University Press, Cambridge, UK.Google Scholar
Berman, S. A., Buggeln, J., Brantley, D. A., Mitchell, K. A., and Solomon, T. H. (2021). Transport barriers to self-propelled particles in fluid flows. Phys. Rev. Fluids, 6, L012501.Google Scholar
Beron-Vera, F. J. (2021). Nonlinear dynamics of inertial particles in the ocean: From drifters and floats to marine debris and Sargassum. Nonlin. Dyn., 103, 126.CrossRefGoogle Scholar
Beron-Vera, F. J., and Miron, P. (2020). A minimal Maxey–Riley model for the drift of Sargassum rafts. J. Fluid Mech., 904, A8.CrossRefGoogle Scholar
Beron-Vera, F. J., Olascoaga, M. J., and Goni, G. J. (2008a). Oceanic mesoscale eddies as revealed by Lagrangian coherent structures. Geophys. Res. Lett., 35(12).Google Scholar
Beron-Vera, F. J., Brown, M. G., Olascoaga, M. J., Rypina, I. I., Koçak, H., and Udovydchenkov, I.A. (2008b). Zonal jets as transport barriers in planetary atmospheres. J. Atmosph. Sci., 65(10), 33163326.Google Scholar
Beron-Vera, F. J., Olascoaga, M. J., Brown, M. G., Koçak, H., and Rypina, I. I. (2010). Invariant-tori-like Lagrangian coherent structures in geophysical flows. Chaos, 20, 017514.CrossRefGoogle ScholarPubMed
Beron-Vera, F. J., Olascoaga, M. J., Brown, M. G., and Koçak, H. (2012). Zonal jets as meridional transport barriers in the subtropical and polar lower stratosphere. J. Atmos. Sci., 69, 753767.Google Scholar
Beron-Vera, F. J., Wang, Y., Olascoaga, M. J., Goni, G. J., and Haller, G. (2013). Objective detection of oceanic eddies and the Agulhas leakage. J. Phys. Oceanogr., 43(7), 14261438.CrossRefGoogle Scholar
Beron-Vera, F. J., Olascoaga, M. J., Haller, G., Farazmand, M., Triñanes, J., and Wang, Y. (2015). Dissipative inertial transport patterns near coherent Lagrangian eddies in the ocean. Chaos, 25(8), 087412.Google Scholar
Beron-Vera, F. J., Olascoaga, M. J., Wang, Y., Triñanes, J., and Pérez-Brunius, P. (2018). Enduring Lagrangian coherence of a Loop Current ring assessed using independent observations. Sci. Rep., 8, 11275.CrossRefGoogle ScholarPubMed
Beron-Vera, F. J., Olascoaga, M. J., and Miron, P. (2019a). Building a Maxey–Riley framework for surface ocean inertial particle dynamics. Phys. Fluids, 31(9), 096602.Google Scholar
Beron-Vera, F. J., Hadjighasem, A., Xia, Q., Olascoaga, M. J., and Haller, G. (2019b). Coherent Lagrangian swirls among submesoscale motions. Proc. Natl. Acad. Sci. USA, 116(37), 1825118256.CrossRefGoogle ScholarPubMed
Bettencourt, J. H., López, C., and Hernández-Garcia, E. (2013). Characterization of coherent structures in three-dimensional turbulent flows using the finite-size Lyapunov exponent. J. Phys. A, 46, 254022.CrossRefGoogle Scholar
Bird, R. B., and Stewart, W. E. (2007). Transport Phenomena. John Wiley and Sons, Inc., New York, NY.Google Scholar
Birkhoff, G. D. (1931). Proof of the ergodic theorem. Proc. Natl. Acad. Sci. USA, 17, 656660.CrossRefGoogle ScholarPubMed
Blazevski, D., and Haller, G. (2014). Hyperbolic and elliptic transport barriers in three-dimensional unsteady flows. Physica D, 273–274, 4662.Google Scholar
Bollt, E. M., and Santitissadeekorn, N. (2013). Applied and Computational Measurable Dynamics. SIAM, Philadephia, PA.Google Scholar
Born, S., Wiebel, A., Friedrich, J., Scheuermann, G., and Bartz, D. (2010). Illustrative stream surfaces. IEEE Trans. Vis. Comp. Graphics, 16(6), 13291338.Google Scholar
Bowman, K. P. (1993). Large-scale isentropic mixing properties of the Antarctic polar vortex from analyzed winds. J. Geophys. Res. Atmos., 98(D12), 2301323027.Google Scholar
Bowman, K. P. (1999). Manifold geometry and mixing in observed atmospheric flows. Unpublished manuscript.Google Scholar
Brach, L., Deixonne, P., Bernard, M.-F., et al. (2018). Anticyclonic eddies increase accumulation of microplastic in the North Atlantic subtropical gyre. Marine Pollution Bull., 126, 191196.CrossRefGoogle ScholarPubMed
Brown, M. G. (1998). Phase space structure and fractal trajectories in 1-1/2 degree of freedom Hamiltonian systems whose time dependence is quasiperiodic. Nonlin. Proc. Geophys., 5(2), 6974.Google Scholar
Brown, M. G., and Samelson, R. M. (1994). Particle motion in vorticity-conserving, two-dimensional incompressible flows. Phys. Fluids, 6(9), 28752876.CrossRefGoogle Scholar
Brunton, S. L., and Rowley, C. W. (2010). Fast computation of finite-time Lyapunov exponent fields for unsteady flows. Chaos, 20, 017503.Google Scholar
Budisić, M., and Mezić, I. (2012). Geometry of the ergodic quotient reveals coherent structures in flows. Physica D, 241(15), 12551269.Google Scholar
Budisić, M., Siegmund, S., Thai Son, D., and Mezić, I. (2016). Mesochronic classification of trajectories in incompressible 3D vector fields over finite times. Disc. Cont. Dyn. Sys., Series S, 9, 923958.Google Scholar
Burns, T. J., Davis, R. W., and Moore, E. F. (1999). A perturbation study of particle dynamics in a plane wake flow. J. Fluid Mech., 384, 126.Google Scholar
Cassel, K. W., and Conlisk, A. T. (2014). Unsteady separation in vortex-induced boundary layers. Phil. Trans. Royal Soc. A., 372(2020), 20130348.Google Scholar
Cencini, M., and Vulpiani, A. (2013). Finite size Lyapunov exponent: review on applications. J. Phys. A, 46(25), 254019.Google Scholar
Chakraborty, P., Balachandar, S., and Adrian, R. (2005). On the relationships between local vortex identification schemes. J. Fluid Mech., 535, 189214.Google Scholar
Chelton, D. B., Gaube, P., Schlax, M. G., Early, J. J., and Samelson, R. M. (2011a). The influence of nonlinear mesoscale eddies on near-surface oceanic chlorophyll. Science, 334(6054), 328332.Google Scholar
Chelton, D. B., Schlax, M. G., and Samelson, R. M. (2011b). Global observations of nonlinear mesoscale eddies. Progr. Oceanogr., 91(2), 167216.Google Scholar
Chen, P. (1994). The permeability of the Antarctic vortex edge. J. Geophys. Res. Atmos., 99(D10), 2056320571.Google Scholar
Chicone, C. (2006). Ordinary Differential Equations with Applications. Springer–Verlag, New York, NY.Google Scholar
Chien, W. L., Rising, H., and Ottino, J. M. (1986). Laminar mixing and chaotic mixing in several cavity flows. J. Fluid Mech., 170, 355377.Google Scholar
Childress, S. (2009). A Theoretical Introduction to Fluid Dynamics. AMS, Providence, RI.Google Scholar
Chong, M. S., Perry, A. E., and Cantwell, B. J. (1990). A general classification of three-dimensional flow field. Phys. Fluids, 2, 765777.Google Scholar
Chorin, A. J., and Marsden, J. E. (1993). A Mathematical Introduction to Fluid Mechanics. Springer, New York, NY.CrossRefGoogle Scholar
Claudel, C.-M., Virbhadra, K. S., and Ellis, G. F. R. (2001). The geometry of photon surfaces. J. Math. Phys., 42(2), 818838.CrossRefGoogle Scholar
Cornfeld, I. P., Fomin, S. V., and Sinai, Ya. G. (1982). Ergodic Theory. Springer, New York, NY.Google Scholar
Cotter, B. A., and Rivlin, R. S. (1955). Tensors associated with time-dependent stress. Q. Appl. Math., 13(5), 177182.CrossRefGoogle Scholar
Coulliette, C., Lekien, F., Paduan, J. D., Haller, G., and Marsden, J. E. (2007). Optimal pollution mitigation in Monterey Bay based on coastal radar data and nonlinear dynamics. Environ. Sci. Technol., 41, 65626572.Google Scholar
Crutzen, P. J. (2006). Albedo enhancement by stratospheric sulfur injections: A contribution to resolve a policy dilemma? Climatic Change, 77, 211219.Google Scholar
Cucitore, R., Quadrio, M., and Baron, A. (1999). On the effectiveness and limitations of local criteria for the identification of a vortex. Eur. J. Mech.B/Fluids, 18, 261282.Google Scholar
Daitche, A. (2013). Advection of inertial particles in the presence of the history force: higher order numerical schemes. J. Comput. Phys., 254, 93106.Google Scholar
Daitche, A., and Tél, T. (2014). Memory effects in chaotic advection of inertial particles. New J. Phys., 16, 073008.Google Scholar
D’Asaro, Eric A., Shcherbina, Andrey Y., Klymak, Jody M., et al. (2018). Ocean convergence and the dispersion of flotsam. Proc. Natl. Acad. Sci. USA, 115(6), 11621167.Google Scholar
De Silva, C. M., Hutchins, N., and Marusic, I. (2015). Uniform momentum zones in turbulent boundary layers. J. Fluid. Mech., 786, 309331.Google Scholar
Dee, D. P., Uppala, S. M., Simmons, Adrian J., et al. (2011). The ERA-Interim reanalysis: Configuration and performance of the data assimilation system. Quart. J. Royal Meteorol. Soc., 137(656), 553597.Google Scholar
del Castillo-Negrete, D., and Morrison, P. J. (1993). Chaotic transport by Rossby waves in shear flow. Phys. Fluids A, 5(4), 948965.CrossRefGoogle Scholar
del Castillo-Negrete, D., Greene, J. M., and Morrison, P. J. (1996). Area preserving nontwist maps: periodic orbits and transition to chaos. Physica D, 91(1), 123.Google Scholar
Délery, J. M. (2001). Robert Legendre and Henri Werlé: Toward the elucidation of three-dimensional separation. Annu. Rev. Fluid Mech., 33, 129154.CrossRefGoogle Scholar
Delmarcelle, T., and Hesselink, L. (1994). The topology of symmetric, second-order tensor fields. Proceedings of the Conference on Visualization’ 94, 140147.Google Scholar
Delshams, A., and de la Llave, R. (2000). KAM theory and a partial justification of Green’s criterion for nontwist maps. SIAM J. Math. Anal., 31(6), 12351269.Google Scholar
Dinklage, A., Klinger, T., Marx, G., and Schweikhard, L. (2005). Plasma Physics – Confinement, Transport and Collective Effects. Springer, Heidelberg, Germany.Google Scholar
Doan, M., Simons, J. J., Lilienthal, K., Solomon, T., and Mitchell, K. A. (2018). Barriers to front propagation in laminar, three-dimensional fluid flows. Phys. Rev. E, 97, 033111.Google Scholar
Dombre, T., Frisch, U., Greene, J. M., et al. (1986). Chaotic streamlines in ABC flows. J. Fluid Mech., 167, 353391.CrossRefGoogle Scholar
Dong, C., McWilliams, J., Liu, Y., and Chen, D. (2014). Global heat and salt transports by eddy movement. Nat. Commun., 5, 3294.Google Scholar
d’Ovidio, F., Fernandez, V., Hernández-Garcia, E., and López, C. (2004). Mixing structures in the Mediterranean Sea from finite-size Lyapunov exponents. Geophys. Res. Lett., 31(17), L17203.Google Scholar
Drouot, R. (1976). Définition d’un transport associé un modèle de fluide de deuxième ordre. C. R. Acad. Sc. Paris, Series A, 282, 923926.Google Scholar
Drouot, R., and Lucius, M. (1976). Approximation du second ordre de la loi de comportement des fluides simples. Lois classiques deduites de l’introduction d’un nouveau tenseur objectif. Archiwum Mechaniki Stosowanej, 28/2, 189198.Google Scholar
Dryden, H. L., von Kármán, T., and Adam, K.A. (1941). Fluid Mechanics and Statistical Methods in Engineering. University of Pennsylvania Press, PA.Google Scholar
Dubief, Y., and Delcayre, F. (2000). On coherent-vortex identification in turbulence. J. Turbulence, 1, N11.CrossRefGoogle Scholar
Eberly, D. (1996). Ridges in Image and Data Analysis. Springer, New York, NY.Google Scholar
Eisma, J., Westerweel, J., Ooms, G., and Elsinga, G. E. (2015). Interfaces and internal layers in a turbulent boundary layer. Phys. Fluids, 27, 055103.CrossRefGoogle Scholar
Elhmaïdi, D., Provenzale, A., and Babiano, A. (1993). Elementary topology of two-dimensional turbulence from a Lagrangian viewpoint and single-particle dispersion. J. Fluid Mech., 257, 533558.Google Scholar
Elipot, S., Lumpkin, R., Perez, R. C., et al. (2016). A global surface drifter data set at hourly resolution. J. Geophys. Res. Oceans, 121(5), 29372966.Google Scholar
Encinas-Bartos, A. P., Aksamit, N., and Haller, G. (2022). Quasi-objective eddy visualization from sparse drifter data. Chaos (submitted) ArXiv: 2111.14117.Google Scholar
Epps, B. (2017). Review of vortex identification methods. AIAA SciTech Forum, 9-13 January, 2017, Grapevine, Texas, 55th AIAA Aerospace Sciences Meeting, 122.Google Scholar
Epstein, I. J. (1963). Conditions for a matrix to commute with its integral. Proc. AMS, 14, 266270.Google Scholar
Ethier, R. C., and Steinman, D. A. (1994). Exact fully 3D Navier–Stokes solutions for benchmarking. Int. J. Numer. Meth. Fluids, 19, 369375.Google Scholar
Everitt, B. S., Landau, S., Leese, M., and Stahl, D. (2011). Cluster Analysis. Wiley, New York, NY.Google Scholar
Falco, R. E. (1977). Coherent motions in the outer region of turbulent boundary layers. Phys. Fluids, 20(10), S124S132.Google Scholar
Fan, D., Xu, J., Yao, M. X., and Hickey, J. P. (2019). On the detection of internal interfacial layers in turbulent flows. J. Fluid Mech., 872, 198217.Google Scholar
Farazmand, M., and Haller, G. (2012). Computing Lagrangian coherent structures from their variational theory. Chaos, 22(1), 013128.Google Scholar
Farazmand, M., and Haller, G. (2013). Attracting and repelling Lagrangian coherent structures from a single computation. Chaos, 23(2), 023101.Google Scholar
Farazmand, M., and Haller, G. (2015). The Maxey–Riley equation: Existence, uniqueness and regularity of solutions. Nonlin. Anal. Real World Applications, 22, 98106.Google Scholar
Farazmand, M., and Haller, G. (2016). Polar rotation angle identifies elliptic islands in unsteady dynamical systems. Physica D, 315, 112.Google Scholar
Farazmand, M., Kevlahan, N. K. R., and Protas, B. (2011). Controlling the dual cascade of two-dimensional turbulence. J. Fluid Mech., 668, 202222.Google Scholar
Farazmand, M., Blazevski, D., and Haller, G. (2014). Shearless transport barriers in unsteady two-dimensional flows and maps. Physica D, 278–279, 4457.Google Scholar
Fenichel, N. (1971). Persistence and smoothness of invariant manifolds for flows. Indiana Univ. Math. J., 21, 193226.Google Scholar
Fenichel, N. (1979). Geometric singular perturbation theory for ordinary differential equations. J. Diff. Eqs., 31(1), 5398.Google Scholar
Fountain, G. O., Khakhar, D. V., Mezić, I., and Ottino, J. M. (2000). Chaotic mixing in a bounded three-dimensional flow. J. Fluid Mech., 417, 265301.Google Scholar
Friedlander, S., and Vishik, M. M. (1992). Instability criteria for steady flows of a perfect fluid. Chaos, 2(3), 455460.Google Scholar
Froyland, G. (2013). An analytic framework for identifying finite-time coherent sets in time-dependent dynamical systems. Physica D, 250, 119.Google Scholar
Froyland, G. (2015). Dynamic isoperimetry and the geometry of Lagrangian coherent structures. Nonlinearity, 28, 35873622.Google Scholar
Froyland, G., and Kwok, E. (2017). A dynamic Laplacian for identifying Lagrangian coherent structures on weighted Riemannian manifolds. J. Nonlin. Sci., 30, 18891971.Google Scholar
Froyland, G., and Padberg-Gehle, K. (2015). A rough-and-ready cluster-based approach for extracting finite-time coherent sets from sparse and incomplete trajectory data. Chaos, 25(8), 087406.Google Scholar
Froyland, G., Santitissadeekorn, N., and Monahan, A. (2010). Transport in time-dependent dynamical systems: Finite-time coherent sets. Chaos, 20, 043116.Google Scholar
Froyland, G., Koltai, P., and Plonka, M. (2020). Computation and optimal perturbation of finite-time coherent sets for aperiodic flows without trajectory integration. SIAM J. Appl. Dynamical Sys., 19, 16591700.Google Scholar
Fyrillas, M. M., and Nomura, K. K. (2007). Diffusion and Brownian motion in Lagrangian coordinates. J. Chem. Phys., 126, 164510.Google Scholar
Gao, F., Ma, W., Zambonini, G., et al. (2015). Large-eddy simulation of 3-D corner separation in a linear compressor cascade. Phys. Fluids, 27, 085105.Google Scholar
Garaboa-Paz, D., Eiras-Barca, J., Huhn, F., and Pérez-Muñuzuri, V. (2015). Lagrangian coherent structures along atmospheric rivers. Chaos, 25(6), 063105.Google Scholar
Gelfand, I. A., and Fomin, S. V. (2000). Calculus of Variations. Dover Publications, Mineola, NY.Google Scholar
Gettelman, A., Hannay, C., Bacmeister, J. T., et al. (2019). High climate sensitivity in the Community Earth System Model Version 2 (CESM2). Geophys. Res. Lett., 46, 83298337.Google Scholar
Ghosh, S., Leonard, A., and Wiggins, S. (1998). Diffusion of a passive scalar from a no-slip boundary into a two-dimensional chaotic advection field. J. Fluid Mech., 372, 119163.Google Scholar
Golé, C. (2001). Symplectic Twist Maps. World Scientific, Singapore.Google Scholar
Golub, G. H., and Van Loan, C. F. (2013). Matrix Computations. Johns Hopkins University Press, Baltimore.Google Scholar
Gowen, S., and Solomon, T. (2015). Experimental studies of coherent structures in an advection-reaction-diffusion system. Chaos, 25(8), 087403.Google Scholar
Graham, M. D., and Floryan, D. (2021). Exact coherent states and the nonlinear dynamics of wall-bounded turbulent flows. Annual Rev. Fluid Mech., 53(1), 227253.Google Scholar
Green, M. A., Rowley, C. W., and Haller, G. (2007). Detection of Lagrangian coherent structures in three-dimensional turbulence. J. Fluid Mech., 572, 111120.Google Scholar
Greene, J. M., and Kim, J.-S. (1987). The calculation of Lyapunov spectra. Physica D, 24(1), 213225.Google Scholar
Gromeka, I. S. (1881). Some cases of incompressible fluid motion. Scientific Notes of the Kazan University, pp. 76148.Google Scholar
Guckenheimer, J., and Holmes, P. (1983). Nonlinear Oscillations, Dynamical Systems and Bifurcation of Vector Fields. Springer, New York, NY.Google Scholar
Guillemin, V., and Pollack, A. (2010). Differential Topology. AMS, Providence, RI.Google Scholar
Günther, T., and Theisel, H. (2018). The state of the art in vortex extraction. Comput. Graph. Forum, 37, 149173.Google Scholar
Günther, T., and Theisel, H. (2020). Hyper-Objective Vortices. IEEE Trans. Vis. Comput. Graph., 26(3), 15321547.Google Scholar
Günther, T., Gross, M., and Theisel, H. (2017). Generic objective vortices for flow visualization. ACM Trans. Graph., 36, 141:111.Google Scholar
Gurtin, M. E. (1981). An Introduction to Continuum Mechanics. Academic Press, New York, NY.Google Scholar
Gurtin, M. E., Fried, E., and Anand, L. (2010). The Mechanics and Thermodynamics of Continua. Cambridge University Press, Cambridge, UK.Google Scholar
Hadjighasem, A., and Haller, G. (2016a). Geodesic transport barriers in Jupiter’s atmosphere: A video-based analysis. SIAM Rev., 6989.Google Scholar
Hadjighasem, A., and Haller, G. (2016b). Level set formulation of two-dimensional Lagrangian vortex detection methods. Chaos, 26, 103102.Google Scholar
Hadjighasem, A., Karrasch, D., Teramoto, H., and Haller, G. (2016). Spectral clustering approach to Lagrangian vortex detection. Phys. Rev. E, 93.Google Scholar
Hadjighasem, A., Farazmand, M., Blazevski, D., Froyland, G., and Haller, G. (2017). A critical comparison of Lagrangian methods for coherent structure detection. Chaos, 27, 053104.Google Scholar
Hadwiger, M., Mlejnek, M., Theussl, T., and Rautek, P. (2019). Time-dependent flow seen through approximate observer killing fields. IEEE Trans. Vis. Comp. Graph., 25, 12571266.Google Scholar
Haley, P. J., and Lermusiaux, P. F. J. (2010). Multiscale two-way embedding schemes for free-surface primitive equations in the “Multidisciplinary Simulation, Estimation and Assimilation System”. Ocean Dynamics, 60, 14971537.Google Scholar
Haller, G. (2000). Finding finite-time invariant manifolds in two-dimensional velocity fields. Chaos, 10, 99108.Google Scholar
Haller, G. (2001a). Distinguished material surfaces and coherent structures in 3D fluid flows. Physica D, 149, 248277.Google Scholar
Haller, G. (2001b). Lagrangian coherent structures and the rate of strain in a partition of two-dimensional turbulence. Phys. Fluids, 13, 33653385.Google Scholar
Haller, G. (2002). Lagrangian coherent structures from approximate velocity data. Phys. Fluids, 14, 18511861.CrossRefGoogle Scholar
Haller, G. (2004). Exact theory of unsteady separation for two-dimensional flows. J. Fluid Mech., 512, 257311.Google Scholar
Haller, G. (2005). An objective definition of a vortex. J. Fluid Mech., 525, 126.CrossRefGoogle Scholar
Haller, G. (2011). A variational theory of hyperbolic Lagrangian coherent structures. Physica D, 240, 574598.Google Scholar
Haller, G. (2015). Lagrangian coherent structures. Ann. Rev. Fluid Mech., 47, 137162.Google Scholar
Haller, G. (2016). Dynamic rotation and stretch tensors from a dynamic polar decomposition. J. Mech. Phys. Solids, 86, 7093.Google Scholar
Haller, G. (2021). Can vortex criteria be objectivized? J. Fluid Mech., 908, A25.CrossRefGoogle Scholar
Haller, G., and Beron-Vera, F. J. (2012). Geodesic theory of transport barriers in two-dimensional flows. Physica D, 241(20), 16801702.Google Scholar
Haller, G., and Beron-Vera, F. J. (2013). Coherent Lagrangian vortices: The black holes of turbulence. J. Fluid Mech., 731, R4.Google Scholar
Haller, G., and Beron-Vera, F. J. (2014). Addendum to “Coherent Lagrangian vortices: The black holes of turbulence”. J. Fluid Mech., 751, R3.Google Scholar
Haller, G., and Iacono, R. (2003). Stretching, alignment, and shear in slowly varying velocity fields. Phys. Rev. E, 68, 056304.Google Scholar
Haller, G., and Mezić, I. (1998). Reduction of three-dimensional, volume-preserving flows with symmetry. Nonlinearity, 11(2), 319339.Google Scholar
Haller, G., and Poje, A. (1998). Finite time transport in aperiodic flows. Physica D, 119, 352380.Google Scholar
Haller, G., and Sapsis, T. (2008). Where do inertial particles go in fluid flows? Physica D, 237, 573583.Google Scholar
Haller, G., and Sapsis, T. (2010). Localized instability and attraction along invariant manifolds. SIAM J. Appl. Dyn. Sys., 9, 611633.Google Scholar
Haller, G., and Sapsis, T. (2011). Lagrangian coherent structures and the smallest finite-time Lyapunov exponent. Chaos, 21(2), 023115.Google Scholar
Haller, G., and Yuan, G. (2000). Lagrangian coherent structures and mixing in two-dimensional turbulence. Physica D, 147, 352370.Google Scholar
Haller, G., Hadjighasem, A., Farazmand, M., and Huhn, F. (2016). Defining coherent vortices objectively from the vorticity. J. Fluid Mech., 795, 136173.Google Scholar
Haller, G., Karrasch, D., and Kogelbauer, F. (2018). Material barriers to diffusive and stochastic transport. Proc. Natl. Acad. Sci. USA, 115, 90749079.Google Scholar
Haller, G., Karrasch, D., and Kogelbauer, F. (2020a). Barriers to the transport of diffusive scalars in compressible flows. SIAM J. Appl. Dyn. Sys., 19(1), 85123.CrossRefGoogle Scholar
Haller, G., Katsanoulis, S., Holzner, B., Frohnapfel, B., and Gatti, D. (2020b). Objective barriers to the transport of dynamically active vector fields. J. Fluid Mech., 905, A17.Google Scholar
Haller, G., Aksamit, N., and Encinas-Bartos, A. P. (2021). Quasi-objective coherent structure diagnostics from single trajectories. Chaos, 31, 043131.Google Scholar
Haller, G., Aksamit, N., and Encinas-Bartos, A. P. (2022). Erratum:“Quasi-objective coherent structure diagnostics from single trajectories” [Chaos 31, 043131 (2021)]. Chaos, 32(5), 059901.Google Scholar
Haza, A. C., D’Asaro, E., Chang, H., et al. (2018). Drogue-loss detection for surface drifters during the Lagrangian Submesoscale Experiment (LASER). J. Atmosph. Ocean. Techn., 35, 705725.Google Scholar
Head, M. R., and Bandyopadhyay, P. (1981). New aspects of turbulent boundary-layer structure. J. Fluid Mech., 107, 297338.Google Scholar
Henderson, K. L., Gwynllyw, D. R., and Barenghi, C. F. (2007). Particle tracking in Taylor–Couette flow. Eur. J. Mech.B/Fluids, 26(6), 738748.Google Scholar
Hill, M. J. M. (1894). On a spherical vortex. Phil. Trans. Royal Soc. A., 185, 213245.Google Scholar
Hua, B. L., and Klein, P. (1998). An exact criterion for the stirring properties of nearly two-dimensional turbulence. Physica D, 113, 98110.Google Scholar
Hua, B. L., McWilliams, J. C., and Klein, P. (1998). An exact criterion for the stirring properties of nearly two-dimensional turbulence. J. Fluid Mech., 366, 87108.Google Scholar
Huhn, F., van Rees, W.M., Gazzola, M., et al. (2015). Quantitative flow analysis of swimming dynamics with coherent Lagrangian vortices. Chaos, 25, 087405.Google Scholar
Hunt, J. C. R., Wray, A., and Moin, P. (1988). Eddies, stream, and convergence zones in turbulent flows. Center for Turb. Res. Rep. CTR-S88, 193208.Google Scholar
Jantzen, R. T., Taira, K., Granlund, K. O., and Ol, M. V. (2014). Vortex dynamics around pitching plates. Phys. Fluids, 26, 065105.Google Scholar
Jeong, J., and Hussain, F. (1995). On the identification of a vortex. J. Fluid Mech., 285, 6994.Google Scholar
Jones, C. K. R. T. (1995). Geometric singular perturbation theory. In Dynamical Systems: Lectures given at the 2nd Session of the Centro Internazionale Matematico Estivo (C.I.M.E.) held in Montecatini Terme, Italy, June 13–22, 1994, 44118.Google Scholar
Jones, C. K. R. T., and Winkler, S. (2002). Invariant Manifolds and Lagrangian Dynamics in the Ocean and Atmosphere. In Fiedler, B. (ed.), Handbook of Dynamical Systems. vol. 2. Elsevier Science, Amsterdam, pp. 5592.Google Scholar
Joseph, B., and Legras, B. (2002). Relation between kinematic boundaries, stirring, and barriers for the Antarctic Polar Vortex. J. Atmosph. Sci., 59(7).Google Scholar
Jung, C., Tél, T., and Ziemniak, E. (1993). Application of scattering chaos to particle transport in a hydrodynamical flow. Chaos, 3, 555568.Google Scholar
Kamphuis, M., Jacobs, G. B., Chen, K., Spedding, G., and Hoeijmakers, H. (2018). Pulse actuation and its effects on separated Lagrangian coherent structures for flow over a cambered airfoil. In Wright, S. D., and Hartsfield, C. R. (eds.), 2018 AIAA Aerospace Sciences Meeting (210059 ed.) https://arc.aiaa.org/doi/10.2514/6.2018-2255.Google Scholar
Karatzas, I., and Shreve, S. (1998). Brownian Motion and Stochastic Calculus. Springer, New York, NY.Google Scholar
Karrasch, D. (2015). Attracting Lagrangian coherent structures on Riemannian manifolds. Chaos, 25(8), 087411.Google Scholar
Karrasch, D., and Haller, G. (2013). Do finite-size Lyapunov exponents detect coherent structures? Chaos, 23, 043126.Google Scholar
Karrasch, D., and Schilling, N. (2020). Fast and robust computation of coherent Lagrangian vortices on very large two-dimensional domains. SMAI J. Comp. Math., 6, 101124.Google Scholar
Karrasch, D., Huhn, F., and Haller, G. (2014). Automated detection of coherent Lagrangian vortices in two-dimensional unsteady flows. Proc. Royal Soc. A, 471, 20140639.Google Scholar
Karrasch, D., Farazmand, M., and Haller, G. (2015). Attraction-based computation of hyperbolic Lagrangian coherent structures. J. Comp. Dynamics, 2, 8393.Google Scholar
Kashimura, H., Abe, M., Watanabe, S., et al. (2017). Shortwave radiative forcing, rapid adjustment, and feedback to the surface by sulfate geoengineering: analysis of the Geoengineering Model Intercomparison Project G4 scenario. Atm. Chem. Phys., 17(5), 33393356.Google Scholar
Kasten, J., Petz, C., Hotz, I., Hege, H. C., and Noack, B. R. (2010). Lagrangian feature extraction of the cylinder wake. Phys. Fluids., 22, 091108.Google Scholar
Kaszás, B., Pedergnana, T., and Haller, G. (2022). The objective deformation component of a velocity field. Eur. J. Mech. B/Fluids, submitted.Google Scholar
Katsanoulis, S. (2020). BarrierTool: Automated extraction of material barriers in two-dimensional velocity fields. https://github.com/haller-group/BarrierTool.Google Scholar
Katsanoulis, S., Farazmand, M., Serra, M., and Haller, G. (2020). Vortex boundaries as barriers to diffusive vorticity transport in two-dimensional flows. Phys. Rev. Fluids, 5, 024701.Google Scholar
Kieburg, M., and Kösters, H. (2016). Exact relation between singular value and eigenvalue statistics. Random Matrices: Theory and Appl., 05(04), 1650015.Google Scholar
Kilic, M. S., Haller, G., and Neishtadt, A. (2005). Unsteady fluid flow separation by the method of averaging. Phys. Fluids, 17(6), 067104.Google Scholar
Kline, S. J., Reynolds, W. C., Schraub, F. A., and Runstadler, P. W. (1967). The structure of turbulent boundary layers. J. Fluid. Mech., 30(4), 741773.Google Scholar
Klonowska-Prosnak, M. E., and Prosnak, W. J. (2001). An exact solution to the problem of creeping flow around circular cylinder rotating in presence of translating plane boundary. Acta Mechanica, 146, 115126.Google Scholar
Klose, B. F., Jacobs, G. B., and Serra, M. (2020a). Kinematics of Lagrangian flow separation in external aerodynamics. AIAA J., 58, 19261938.Google Scholar
Klose, B. F., Jacobs, G. B., and Serra, M. (2020b). Objective early identification of kinematic instabilities in shear flows. ArXiv: 2009.05851.Google Scholar
Knobloch, E., and Merryfield, W. J. (1992). Enhancement of diffusive transport in oscillatory flows. Astrophys. J., 401, 196205.Google Scholar
Knobloch, E., and Weiss, J. B. (1987). Chaotic advection by modulated traveling waves. Phys. Rev. A, 36, 15221524.Google Scholar
Kolář, V. (2007). Vortex identification: New requirements and limitations. Int. J. Heat Fluid Flow, 28(4), 638652.Google Scholar
Kravitz, B., MaMartin, D. G., Mills, M. J., et al. (2017). First simulations of designing stratospheric sulfate aerosol geoengineering to meet multiple simultaneous climate objectives. J. Geophys. Res. Atmos., 122, 12,61612,634.Google Scholar
Kulkarni, C. (2021). Prediction, Analysis and Learning of Advective Transport in Dynamic Fluid Flows. Ph.D. Thesis, Massachusetts Institute of Technology, Department of Mechanical Engineering.Google Scholar
LaCasce, J. H. (2008). Statistics from Lagrangian observations. Progr. Oceanogr., 77, 129.Google Scholar
Lai, Y.-C., and Tél, T. (2011). Transient Chaos. Springer, New York.Google Scholar
Landau, L. D., and Lifshitz, E. M. (1966). Fluid Mechanics. Pergamon Press.Google Scholar
Langlois, G. P., Farazmand, M., and Haller, G. (2015). Asymptotic dynamics of inertial particles with memory. J. Nonlin. Sci, 25, 12251255.Google Scholar
Lapeyre, G., Klein, P., and Hua, B. L. (1999). Does the tracer gradient vector align with the strain eigenvectors in 2-D turbulence? Phys. Fluids, 11, 37293737.Google Scholar
Lapeyre, G., Hua, B. L., and Klein, P. (2001). Dynamics of the orientation of active and passive scalars in two-dimensional turbulence. Phys. Fluids, 13, 251264.Google Scholar
Lebreton, L., Slat, B., Ferrari, F., et al. (2018). Evidence that the Great Pacific Garbage Patch is rapidly accumulating plastic. Sci. Rep., 8, 4666.Google Scholar
Lekien, F., and Haller, G. (2008). Unsteady flow separation on slip boundaries. Phys. Fluids, 20(9), 097101.Google Scholar
Lekien, F., and Ross, S. D. (2010). The computation of finite-time Lyapunov exponents on unstructured meshes and for non-Euclidean manifolds. Chaos, 20, 017505.Google Scholar
Lekien, F., Coulliette, C., Mariano, A. J., et al. (2005). Pollution release tied to invariant manifolds: A case study for the coast of Florida. Physica D, 210, 120.Google Scholar
Lekien, F., Shadden, S. C., and Marsden, J. E. (2007). Lagrangian coherent structures in n-dimensional systems. J. Math. Phys., 48(6), 065404.Google Scholar
Lewis, J. P. (1969). Homogeneous functions and Euler’s theorem. In An Introduction to Mathematics. Macmillan, London.Google Scholar
Li, Y., Perlman, E., Wan, M., et al. (2008). A public turbulence database cluster and applications to study Lagrangian evolution of velocity increments in turbulence. J. Turbulence, 9, N31.Google Scholar
Lighthill, M. J. (1963). Introduction: Boundary layer theory. In Rosenhead, L. (ed), Laminar Boundary Layers. Oxford University Press, Oxford, pp. 46113.Google Scholar
Lim, T. T., and Nickels, T. B. (1995). Vortex Rings. In Green, S.I. (ed.) Fluid Vortices, pp. 95153. Springer, Dordrecht, Netherlands.Google Scholar
Limaye, S. S. (1986). New estimates of the mean zonal flow at the cloud level. Icarus, 65, 335352.Google Scholar
Liu, J., Gao, Y., Wang, Y, and Liu, C. (2019a). Objective Omega vortex identification method. J. Hydrodynam., 31, 455463.Google Scholar
Liu, J., Gao, Y., and Liu, C. (2019b). An objective version of the Rortex vector for vortex identification. Phys. Fluids, 31(6), 065112.Google Scholar
Liu, T., Abernathey, R., Sinha, A., and Chen, D. (2019). Quantifying Eulerian eddy leakiness in an idealized model. J. Geophys. Res. Oceans, 124(12), 88698886.Google Scholar
Liu, W., and Haller, G. (2004). Strange eigenmodes and decay of variance in the mixing of diffusive tracers. Physica D, 188, 139.Google Scholar
Llibre, J., and Valls, C. (2012). A note on the first integrals of the ABC system. J. Math. Phys., 53(2), 023505.Google Scholar
Lugt, H. J. (1979). The dilemma of defining a vortex. In Muller, U., Riesner, K. G., and Schmidt, B. (eds), Recent Developments in Theoretical and Experimental Fluid Mechanics. pp. 13, 309321.Google Scholar
Lumpkin, R., and Pazos, M. (2007). Lagrangian Analysis and Prediction in Coastal and Ocean Processes. Cambridge University Press, Cambridge, UK.Google Scholar
Ma, T., and Bollt, E. M. (2013). Relatively coherent sets as a hierarchical partition method. Int. J. Bifurc. Chaos, 23(07), 1330026.Google Scholar
Mackay, R. S. (1994). Transport in 3D volume-preserving flows. J. Nonlin. Sci., 4, 329354.Google Scholar
Mackay, R. S., Meiss, J. D., and Percival, I. C. (1984). Transport in Hamiltonian systems. Physica D, 13, 5581.Google Scholar
Madrid, J. A. J., and Mancho, A. M. (2009). Distinguished trajectories in time dependent vector fields. Chaos, 19, 013111.Google Scholar
Mahoney, J. R., and Mitchell, K. A. (2015). Finite-time barriers to front propagation in two-dimensional fluid flows. Chaos, 25, 087404.Google Scholar
Mahoney, J. R., Bargteil, D., Kingsbury, M., Mitchell, K., and Solomon, T. (2012). Invariant barriers to reactive front propagation in fluid flows. Europhys. Lett., 98, 44005.Google Scholar
Majda, A. J., and Bertozzi, A. L. (2002). Vorticity and Incompressible Flow. Cambridge University Press, Cambridge, UK.Google Scholar
Malhotra, N., Mezić, I., and Wiggins, S. (1998). Patchiness: A new diagnostic for Lagrangian trajectory analysis in time-dependent fluid flows. Int. J. Bifurc. Chaos, 08(06), 10531093.Google Scholar
Mancho, A. M., Wiggins, S., Curbelo, J., and Mendoza, C. (2013). Lagrangian descriptors: A method for revealing phase space structures of general time dependent dynamical systems. Comm. Nonlin. Sci. Num. Sim., 18, 35303557.Google Scholar
Mañe, R. (1978). Persistent manifolds are normally hyperbolic. Trans. Am. Math. Soc., 21, 261283.Google Scholar
Martins, R. S., Pereira, A. S., Mompean, G., Thais, L., and Thompson, R. L. (2016). An objective perspective for classic flow classification criteria. C. R. Mécanique, 344, 5259.Google Scholar
Mathur, M., Haller, G., Peacock, T., Ruppert-Felsot, J. E., and Swinney, H. L. (2007). Uncovering the Lagrangian skeleton of turbulence. Phys. Rev. Lett., 98, 144502.Google Scholar
Maxey, M. R. (1987). The gravitational settling of aerosol particles in homogeneous turbulence and random flow fields. J. Fluid Mech., 174, 441465.Google Scholar
Maxey, M. R., and Riley, J. J. (1983). Equation of motion for a small rigid sphere in a nonuniform flow. Phys. Fluids, 26, 883889.Google Scholar
McMullan, W. A., and Page, G. J. (2012). Towards large eddy simulation of gas turbine compressors. Progr. Aerospace. Sci., 52, 3047.Google Scholar
Meiss, J. D. (1992). Symplectic maps, variational principles and transport. Rev. Modern Phys,, 64, 795848.Google Scholar
Mendoza, C., and Mancho, A. M. (2010). Hidden geometry of ocean flows. Phys. Rev. Lett., 105, 038501.Google Scholar
Meunier, P., Le Dizès, S., and Leweke, T. (2005). Physics of vortex merging. Comptes Rendus Physique, 6(4), 431450.Google Scholar
Meyers, J., and Meneveau, C. (2013). Flow visualization using momentum and energy transport tubes and applications to turbulent flow in wind farms. J. Fluid Mech., 715, 335358.Google Scholar
Mezić, I., and Sotiropoulos, F. (2002). Ergodic theory and experimental visualization of invariant sets in chaotically advected flows. Phys. Fluids, 14(7), 22352243.Google Scholar
Mezić, I., Loire, S., Fonoberov, V. A., and Hogan, P. (2010). A new mixing diagnostic and Gulf oil spill movement. Science, 330, 486489.Google Scholar
Michaelides, E. E. (1997). The transient equation of motion for particles, bubbles, and droplets. J. Fluids Eng., 119, 223247.Google Scholar
Miron, P., and Vétel, J. (2015). Towards the detection of moving separation in unsteady flows. J. Fluid Mech., 779, 8184.Google Scholar
Miron, P., Olascoaga, M. J., Beron-Vera, , et al. (2020). Clustering of marine debris and sargassum-like drifters explained by inertial particle dynamics. Geophys. Res. Lett., 47(19), e2020GL089874.Google Scholar
Mitchell, K. A., and Mahoney, J. R. (2012). Invariant manifolds and the geometry of front propagation in fluid flows. Chaos, 22(3), 037104.Google Scholar
Mograbi, E., and Bar-Ziv, E. (2006). On the asymptotic solution of the Maxey–Riley equation. Phys. Fluids, 18(5), 051704.Google Scholar
Monin, A. S., and Yaglom, A. M. (2007). Statistical Fluid Mechanics: Mechanics of Turbulence. Volume I. Dover Publications, Mineola, NY.Google Scholar
Nakamura, N. (2008). Quantifying inhomogeneous, instantaneous, irreversible transport using passive tracer field as a coordinate. Lect. Notes in Phys., 744, 137144.Google Scholar
Neamtu-Halic, M. M., Krug, D., Haller, G., and Holzner, M. (2019). Lagrangian coherent structures and entrainment near the turbulent/non-turbulent interface of a gravity current. J. Fluid Mech., 877, 824843.Google Scholar
Neamtu-Halic, M. M., Krug, D., Mollicone, J.-P., et al. (2020). Connecting the time evolution of the turbulence interface to coherent structures. J. Fluid Mech., 898, A3.Google Scholar
Neff, P., Lankeit, J., and Madeo, A. (2014). On Grioli’s minimum property and its relation to Cauchy’s polar decomposition. Int. J. Eng. Sci., 80, 209217.Google Scholar
Nelson, D. A., and Jacobs, G. B. (2015). DG-FTLE: Lagrangian coherent structures with high-order discontinuous-Galerkin methods. J. Comp. Phys., 295, 6586.Google Scholar
Nolan, P. J., Serra, M., and Ross, S. D. (2020). Finite-time Lyapunov exponents in the instantaneous limit and material transport. Nonlin. Dyn., 100(19), 38253852.Google Scholar
Norgard, G., and Bremer, P.-T. (2012). Second derivative ridges are straight lines and the implications for computing Lagrangian coherent structures. Physica D, 241, 14751476.Google Scholar
Oberlack, M., and Cheviakov, A. F. (2010). Higher-order symmetries and conservation laws of the G-equation for premixed combustion and resulting numerical schemes. J. Eng. Math., 66, 121140.Google Scholar
Oettinger, D. (2017). Variational Approach to Lagrangian Coherent Structures in Three-Dimensional Unsteady Flows. Ph.D. Thesis, ETH Zurich.Google Scholar
Oettinger, D., and Haller, G. (2016). An autonomous dynamical system captures all LCSs in three-dimensional unsteady flows. Chaos, 26(10), 103111.Google Scholar
Oettinger, D., Blazevski, D., and Haller, G. (2016). Global variational approach to elliptic transport barriers in three dimensions. Chaos, 26(3), 033114.Google Scholar
Oettinger, D., Ault, J. T., Stone, H. A., and Haller, G. (2018). Invisible anchors trap particles in branching junctions. Phys. Rev. Lett., 121, 054502.Google Scholar
Ogden, R. W. (1984). Non-Linear Elastic Deformations. Ellis Horwood, Chichester.Google Scholar
Okubo, A. (1970). Horizontal dispersion of floatable trajectories in the vicinity of velocity singularities such as convergencies. Deep-Sea Res., 17, 445454.Google Scholar
Onu, K., Huhn, F., and Haller, G. (2015). LCSTool: A computational platform for Lagrangian coherent structures. J. Comp. Sci., 7, 2636.Google Scholar
Oseledec, V. I. (1968). A multiplicative ergodic theorem. Lyapunov characteristic numbers for dynamical systems. Trudy Moskov. Mat. Obsc, 19, 179210.Google Scholar
Ott, W., and Yorke, J. A. (2008). When Lyapunov exponents fail to exist. Phys. Rev. E, 78, 056203.Google Scholar
Ottino, J. M. (1989). The Kinematics of Mixing: Stretching, Chaos and Transport. Cambridge University Press, Cambridge, UK.Google Scholar
Paduan, J. D., and Cook, M. S. (1997). Mapping surface currents in Monterey Bay with CODAR-type HR radar. Oceanogr., 10, 4952.Google Scholar
Palis, P. Jr. (1969). On Morse–Smale dynamical systems. Topology, 8, 385404.Google Scholar
Palmerius, K. L., Cooper, M., and Ynnerman, A. (2009). Flow field visualization using vector field perpendicular surfaces. Proceedings of the 25th Spring Conference on Computer Graphics. SCCG ’09. Association for Computing Machinery, New York, NY, pp. 2734.Google Scholar
Pedergnana, T., Oettinger, D., and Haller, G. (2020). Explicit unsteady Navier–Stokes solutions and their analysis via local vortex criteria. Phys. Fluids, 32, 046603.Google Scholar
Peikert, R., and Sadlo, F. (2009). Topologically relevant stream surfaces for flow visualization. Proc. of the 25th Spring Conference on Computer Graphics. SCCG ’09. Association for Computing Machinery, New York, NY, pp. 3542.Google Scholar
Peng, J., and Dabiri, J. O. (2009). Transport of inertial particles by Lagrangian coherent structures: application to predator-prey interaction in jellyfish feeding. J. Fluid Mech., 623, 7584.Google Scholar
Perry, A. E., and Chong, M. S. (1987). A description of eddying motions and flow patterns using critical-point concepts. Annual Rev. Fluid Mech., 19, 125155.Google Scholar
Perry, A. E., and Chong, M. S. (1994). Topology of flow patterns in vortex motions and turbulence. Applied. Sci. Res., 53, 357374.Google Scholar
Pierce, J. R., Winstein, D. K., Heckendorn, P., Peter, T., and Keith, D. W. (2010). Efficient formation of stratospheric aerosol for climate engineering by emission of condensible vapor from aircraft. Geophys. Res. Lett., 37, L18805.Google Scholar
Pierrehumbert, R. T. (1991). Large-scale horizontal mixing in planetary atmospheres. Phys. Fluids A, 3(5), 1251260.Google Scholar
Pierrehumbert, R. T. (1994). Tracer microstructure in the large-eddy dominated regime. Chaos, Solit. Fractals, 4, 10911110.Google Scholar
Pierrehumbert, R. T., and Yang, H. (1993). Global chaotic mixing on isentropic surfaces. J. Atmos. Sci., 50, 24622480.Google Scholar
Poje, A. C., Özgökmen, T. M., Lipphardt, B. L., et al. (2014). Submesoscale dispersion in the vicinity of the Deepwater Horizon spill. Proc. Natl. Acad. Sci. USA, 111(35), 1269312698.Google Scholar
Prandtl, L. (1904). Über Flüssigkeitsbewegung bei sehr kleiner Reibung. In Verh. III, Int. Math. Kongr., Heidelberg, 484491.Google Scholar
Prasath, S. G., Vasan, V., and Govindarajan, R. (2019). Accurate solution method for the Maxey–Riley equation, and the effects of Basset history. J. Fluid Mech., 868, 428460.Google Scholar
Pratt, L., Barkan, R., and Rypina, I. (2016). Scalar flux kinematics. Fluids, 1, 27.Google Scholar
Press, W. H., and Rybicki, G. B. (1981). Enhancement of passive diffusion and suppression of heat flux in a fluid with time-varying shear. Astrophys. J., 248, 751766.Google Scholar
Provenzale, A. (1999). Transport by coherent barotropic vortices. Annu. Rev. Fluid Mech., 31, 5593.Google Scholar
Rautek, P., Mlejnek, M., Beyer, J., et al. (2021). Objective observer-relative flow visualization in curved spaces for unsteady 2D geophysical flows. IEEE Trans. Vis. Comp. Graphics, 27, 283293.Google Scholar
Risken, H. (1984). The Fokker–Planck Equation: Methods of Solution and Applications. Springer, New York, NY.Google Scholar
Rojo, I. B., and Günther, T. (2020). Vector field topology of time-dependent flows in a steady reference frame. IEEE Trans. Vis. Comp. Graphics, 26, 12801290.Google Scholar
Rom-Kedar, V. (1994). Homoclinic tangles: classification and applications. Nonlinearity, 7, 441473.Google Scholar
Rom-Kedar, V., and Wiggins, S. (1990). Transport in two-dimensional maps. Arch. Rat. Mech. Anal., 109, 239298.Google Scholar
Rom-Kedar, V., Leonard, A., and Wiggins, S. (1990). An analytical study of transport, mixing and chaos in an unsteady vortical flow. J. Fluid Mech., 214, 347394.Google Scholar
Rosner, D. (2000). Transport Processes in Chemically Reacting Flow Systems. Dover Publications, Mineola, NY.Google Scholar
Rothstein, D., Henry, E., and Gollub, J. P. (1999). Persistent patterns in transient chaotic fluid mixing. Nature, 401, 770772.Google Scholar
Rouche, M., Habets, P., and Laloy, M. (1977). Stability Theory by Lyapunov’s Direct Method. Springer–Verlag New York, NY.Google Scholar
Ruban, A. I., Araki, D., Yapalparvi, R., and Gajjar, J. S. B. (2011). On unsteady boundary-layer separation in supersonic flow. Part 1. Upstream moving separation point. J. Fluid Mech., 678, 124155.Google Scholar
Rubin, J., Jones, C. K. R. T., and Maxey, M. (1995). Settling and asymptotic motion of aerosol particles in a cellular flow field. J. Nonlin. Sci., 5, 337358.Google Scholar
Ruiz-Herrera, A. (2015). Some examples related to the method of Lagrangian descriptors. Chaos, 25, 063112.Google Scholar
Ruiz-Herrera, A. (2016). Performance of Lagrangian descriptors and their variants in incompressible flows. Chaos, 26, 103116.Google Scholar
Rypina, I. I., Brown, M. G., Beron-Vera, F. J., et al. (2007). Area preserving nontwist maps: periodic orbits and transition to chaos. Phys. Rev. Lett., 98, 104102.Google Scholar
Rypina, I. I., Scott, S. E., Pratt, L. J., and Brown, M. G. (2011). Investigating the connection between complexity of isolated trajectories and Lagrangian coherent structures. Nonlin. Proc. Geophys., 18(6), 977987.Google Scholar
Sabelfeld, K. K., and Simonov, N. A. (2012). Random Fields and Stochastic Lagrangian Models. De Gruyter, Berlin, Germany.Google Scholar
Saffman, P. G., (1992). Vortex Dynamics. Cambridge University Press, Cambridge, UK.Google Scholar
Sahner, J., Weinkauf, T., Teuber, N., and Hege, H. (2007). Vortex and strain skeletons in Eulerian and Lagrangian frames. IEEE Trans. Vis. Comp. Graphics, 13(5), 980990.Google Scholar
Samelson, R. M. (1992). Fluid exchange across a meandering jet. J. Phys. Oceanogr., 22, 431440.Google Scholar
Samelson, R. M., and Wiggins, S. (2006). Lagrangian Transport in Geophysical Jets and Waves. Springer-Verlag, New York, NY.Google Scholar
Sanders, J. A., Verhulst, F., and Murdock, J. (2007). Averaging Methods in Nonlinear Dynamical Systems. Springer, New York, NY.Google Scholar
Sane, S., Bujack, R., Garth, C., and Childs, H. (2020). A survey of seed placement and streamline selection techniques. Comput. Graph. Forum, 39(3), 785809.Google Scholar
Sapsis, T., and Haller, G. (2008). Instabilities in the dynamics of neutrally buoyant particles. Phys. Fluids, 20, 017102.Google Scholar
Sapsis, T., and Haller, G. (2009). Inertial particle dynamics in a hurricane. J. Atmosph. Sci., 66(8), 24812492.Google Scholar
Sapsis, T., and Haller, G. (2010). Clustering criterion for inertial particles in two-dimensional time-periodic and three-dimensional steady flows. Chaos, 20, 017515.Google Scholar
Sapsis, T., Peng, J., and Haller, G. (2011a). Instabilities of prey dynamics in jellyfish feeding. Bull. Math. Biol., 73, 18411856.Google Scholar
Sapsis, T. P., Ouellette, N. T., Gollub, J. P., and Haller, G. (2011b). Neutrally buoyant particle dynamics in fluid flows: Comparison of experiments with Lagrangian stochastic models. Phys. Fluids, 23, 093304.Google Scholar
Schindler, B., Peikert, R., Fuchs, R., and Theisel, H. (2012). Ridge concepts for the visualization of Lagrangian coherent structures. In Peikert, R., Hauser, H., Carr, H., and Fuchs, R. (eds), Topological Methods in Data Analysis and Visualization II, pp. 221236.Google Scholar
Schlueter-Kuck, K. L., and Dabiri, J. O. (2017). Coherent structure colouring: identification of coherent structures from sparse data using graph theory. J. Fluid Mech., 811, 468486.Google Scholar
Sears, W. R., and Telionis, D. P. (1975). Boundary-layer separation in unsteady slow. SIAM J. Appl. Math., 28(1), 215235.Google Scholar
Serra, M., and Haller, G. (2016). Objective Eulerian coherent structures. Chaos, 26, 053110.Google Scholar
Serra, M., and Haller, G. (2017a). Efficient computation of null-geodesics with applications to coherent vortex detection. Proc. Royal Soc. A, 473, 2016080.Google Scholar
Serra, M., and Haller, G. (2017b). Forecasting long-lived Lagrangian vortices from their objective Eulerian footprints. J. Fluid Mech., 813, 436457.Google Scholar
Serra, M., Sathe, P., Beron-Vera, F., and Haller, G. (2017). Uncovering the edge of the polar vortex. J. Atmosph. Sci., 74(11), 38713885.Google Scholar
Serra, M., Vétel, J., and Haller, G. (2018). Exact theory of material spike formation in flow separation. J. Fluid Mech., 845, 5192.Google Scholar
Serra, M., Crouzat, S., Simon, G., Vétel, J., and Haller, G. (2020a). Material spike formation in highly unsteady separated flows. J. Fluid Mech., 883, A30.Google Scholar
Serra, M., Sathe, P., Rypina, I., et al. (2020b). Search and rescue at sea aided by hidden flow structures. Nat. Commun., 11, 2525.Google Scholar
Shadden, S. C. (2011). Lagrangian coherent structures. In Grigoriev, R. (ed.), Transport and Mixing in Laminar Flows: From Microfluidics to Oceanic Currents, Wiley-VCH, Berlin, Germany, pp. 5989.Google Scholar
Shadden, S. C., Lekien, F., and Marsden, J. E. (2005). Definition and properties of Lagrangian coherent structures from finite-time Lyapunov exponents in two-dimensional aperiodic flows. Physica D, 212(3), 271304.Google Scholar
Shapiro, A. (1961). Vorticity. Part 1. US National Committee for Fluid Mechanics Film Series. MIT, Cambridge, MA. http://web.mit.edu/hml/ncfmf.html Google Scholar
Shariff, K., Pulliam, T. H., and Ottino, J. M. (1991). A dynamical systems analysis of kinematics in the time-periodic wake of a circular cylinder. Lect. Appl. Math., 28, 613646.Google Scholar
Shariff, K., Leonard, A., and Ferziger, J. H. (2006). Dynamical systems analysis of fluid transport in time-periodic vortex ring flows. Phys. Fluids, 18(4), 047104.Google Scholar
Shepherd, T. G., Koshyk, J. N., and Ngan, K. (2000). On the nature of large-scale mixing in the stratosphere and mesosphere. J. Geophys. Res., 105(D10), 1243312446.Google Scholar
Simon-Miller, A. A., Rogers, J. H., Gierasch, P. J., et al. (2012). Longitudinal variation and waves in Jupiter’s south equatorial wind jet. Icarus, 218, 817830.Google Scholar
Simpson, R. L. (1996). Aspects of turbulent boundary layer separation. Prog. Aerospace Sci., 32, 457521.Google Scholar
Smale, S. (1967). Differentiable dynamical systems. Bull. AMS, 73, 747817.Google Scholar
Smith, F. T. (1986). Steady and unsteady boundary-layer separation. Annu. Rev. Fluid Mech., 18(1), 197220.CrossRefGoogle Scholar
Sotiropoulos, F., Ventikos, Y., and Lackey, T. C. (2001). Chaotic advection in three-dimensional stationary vortex-breakdown bubbles: Šil’nikov’s chaos and the devil’s staircase. J. Fluid Mech., 444, 257297.Google Scholar
Sotiropoulos, F., Webster, D. L., and Lackey, T. C. (2002). Experiments on Lagrangian transport in steady vortex-breakdown bubbles in a confined swirling flow. J. Fluid Mech., 466, 215248.Google Scholar
Speetjens, M., Metcalfe, G., and Rudman, M. (2021). Lagrangian transport and chaotic advection in three-dimensional laminar flows. Appl. Mech. Rev., 73(3), 030801.Google Scholar
Spivak, M. (1999). A Comprehensive Introduction to Differential Geometry, vol. 3, 3rd edn. Publish or Perish, Inc., Houston, TX.Google Scholar
Stevenson, A. F. (1954). Note on the existence and determination of a vector potential. Quart. J. Appl. Math., 12, 194198.Google Scholar
Surana, A., and Haller, G. (2008). Ghost manifolds in slow-fast systems, with application to unsteady fluid flow separation. Physica D, 237(10–12), 15071529.Google Scholar
Surana, A., Grunberg, O., and Haller, G. (2006). Exact theory of three-dimensional flow separation. Part I. Steady separation. J. Fluid Mech., 564, 57103.Google Scholar
Surana, A., Jacobs, G. B., and Haller, G. (2007). Extraction of separation and attachment surfaces from three-dimensional steady shear flows. AIAA J., 45(6), 12901302.Google Scholar
Surana, A., Jacobs, G., Grunberg, O., and Haller, G. (2008). An exact theory of three-dimensional fixed separation in unsteady flows. Phys. Fluids, 20, 107101.Google Scholar
Tabor, M., and Klapper, I. (1994). Stretching and alignment in chaotic and turbulent flows. Chaos, Solitons Fract., 4, 10311055.Google Scholar
Tala, T., and Garbet, X. (2006). Physics of internal transport barriers. Comptes Rendus Physique, 7(6), 622633.Google Scholar
Tang, W., Haller, G., Baik, J.-J., and Ryu, Y.-H. (2009). Locating an atmospheric contamination source using slow manifolds. Phys. Fluids, 21, 043302.Google Scholar
Tang, W., Chan, P. W., and Haller, G. (2010). Accurate extraction of Lagrangian coherent structures over finite domains with application to flight data analysis over Hong Kong International Airport. Chaos, 20, 017502.Google Scholar
Tang, W., Chan, P. W., and Haller, G. (2011a). Lagrangian coherent structure analysis of terminal winds detected by LIDAR. Part I: Turbulence structures. J. Appl. Meteorol. Climatol., 50(2), 325338.Google Scholar
Tang, W., Chan, P. W., and Haller, G. (2011b). Lagrangian coherent structure analysis of terminal winds detected by LIDAR. Part II: Structure evolution and comparison with flight data. J. Appl. Meteorol. Climatol., 50(10), 21672183.Google Scholar
Tang, X. Z., and Boozer, A. H. (1996). Finite time Lyapunov exponent and advection-diffusion equation. Physica D, 95, 283305.Google Scholar
Tanga, P., Babiano, A., Dubrulle, B., and Provenzale, A. (1996). Forming planetesimals in vortices. Icarus, 121(1), 158170.Google Scholar
Tél, T., de Moura, A., Grebogi, C., and Károlyi, G. (2005). Chemical and biological activity in open flows: A dynamical system approach. Phys. Rep., 413(2), 91196.Google Scholar
Theisel, H., Hadwiger, M., Rautek, P., Theussl, T., and Günther, T. (2021). Vortex criteria can be objectivized by unsteadiness minimization. Phys. Fluids, 33(10), 107115.Google Scholar
Theisel, H., Friederici, A., and Günther, T. (2022). Objective flow measures based on few trajectories. ArXiv: 2202.09566.Google Scholar
Thiffeault, J.-L. (2003). Advection-diffusion in Lagrangian coordinates. Phys. Lett. A, 30, 415422.Google Scholar
Thiffeault, J.-L. (2008). Scalar decay in chaotic mixing. Lect. Notes Phys., 744, 335.Google Scholar
Thiffeault, J.-L., Gouillart, E., and Dauchot, O. (2011). Moving walls accelerate mixing. Phys. Rev. E, 84, 036313.Google Scholar
Tian, S., Gao, Y., Dong, X., and Liu, C. (2018). Definitions of vortex vector and vortex. J. Fluid Mech., 849, 312339.Google Scholar
Tobak, M., and Peake, D. J. (1982). Topology of three-dimensional separated flows. Annu. Rev. Fluid Mech., 14, 6185.Google Scholar
Toda, M. (2005). Global aspects of chemical reactions in multidimensional phase space. In Toda, M., Komatsuzaki, T., Konishi, T., Berry, R. S., and Rice, S. A. (eds), Geometrical Structures of Phase Space In Multi-Dimensional Chaos: Applications To Chemical Reaction Dynamics In Complex Systems. John Wiley & Sons, New York, NY.Google Scholar
Tran-Cong, T. (1990). On the potential of a solenoidal vector field. J. Math. Anal. Appl., 151, 557580.Google Scholar
Trefethen, L. N., and Bau, D. (1997). Numerical Linear Algebra. SIAM, Philadelphia, PA.Google Scholar
Truesdell, C. A. (1992). A First Course in Rational Continuum Mechanics. Academic Press, New York, NY.Google Scholar
Truesdell, C., and Rajagopal, K. R. (1999). An Introduction to the Mechanics of Fluids. Birkhäuser, Boston, MA.Google Scholar
Urban, O., Kurková, M., and Rudolf, P. (2021). Application of computer graphics flow visualization methods in vortex rope investigations. Energies., 14(3), 623.Google Scholar
Van Dommelen, L. L. (1981). Unsteady Boundary Layer Separation. Ph.D. Thesis, Cornell University.Google Scholar
Van Dommelen, L. L., and Cowley, S. J. (1990). On the Lagrangian description of unsteady boundary-layer separation. Part 1. General theory. J. Fluid Mech., 210, 593626.Google Scholar
Van Dommelen, L. L., and Shen, S. F. (1982). The genesis of separation. In Cebeci, Tuncer (ed.), Numerical and Physical Aspects of Aerodynamic Flows. Springer, Berlin, Germany, pp. 293311.Google Scholar
Van Dyke, M. (1982). An Album of Fluid Motion. The Parabolic Press, Stanford, CA.Google Scholar
Verhulst, F. (2000). Nonlinear Differential Equations and Dynamical Systems. Springer-Verlag, Berlin, Germany.Google Scholar
Verhulst, F. (2005). Methods and Applications of Singular Perturbations. Springer, New York, NY.Google Scholar
Viana, R. L., Caldas, I. L., Szezech, J. D. Jr., et al. (2021). Transport barriers in symplectic maps. Braz. J. Phys., 51, 899909.Google Scholar
Villermaux, E., and Duplat, J. (2003). Mixing is an aggregation process. C. R. Mécanique, 331(7), 515523.Google Scholar
Voth, G. A., Haller, G., and Gollub, J. P. (1994). Experimental measurements of stretching fields in fluid mixing. Phys. Rev. Lett., 88(25), 254501.Google Scholar
Waleffe, F. (1998). Three-dimensional coherent states in plane shear flows. Phys. Rev. Lett., 81, 41404143.Google Scholar
Waleffe, F. (2001). Exact coherent structures in channel flow. J. Fluid Mech., 435, 93102.Google Scholar
Walters, P. (1982). An Introduction to Ergodic Theory. Springer-Verlag, New York, NY.Google Scholar
Wan, Z. Y., and Sapsis, T. P. (2018). Machine learning the kinematics of spherical particles in fluid flows. J. Fluid Mech., 857, R2.Google Scholar
Wang, K. C. (1972). Separation patterns of boundary layer over an inclined body of revolution. AIAA J., 10, 10441050.Google Scholar
Wang, K. C. (1974). Boundary layer over a blunt body at high incidence with an open-type separation. Proc. Royal Soc. Lond. A, 340, 3355.Google Scholar
Wang, Y., Haller, G., Banaszuk, A., and Tadmor, G. (2003). Closed-loop Lagrangian separation control in a bluff body shear flow model. Phys. Fluids, 15(8), 22512266.Google Scholar
Weiss, J. (1991). The dynamics of entrophy transfer in two-dimensional hydrodynamics. Physica D, 48, 273294.Google Scholar
Weiss, J. B., and Provenzale, A. (2008). Transport and Mixing in Geophysical Flows. Springer, Berlin, Germany.Google Scholar
Weldon, M., Peacock, T., Jacobs, G. B., Helu, M., and Haller, G. (2008). Experimental and numerical investigation of the kinematic theory of unsteady separation. J. Fluid Mech., 611, 111.Google Scholar
Westerweel, J., Fukoshima, C., Pedersen, J. M., and Hunt, J. C. R. (2009). Momentum and scalar transport at the turbulent/non-turbulent interface of a jet. J. Fluid Mech., 631, 199230.Google Scholar
Wiggins, S. (1992). Chaotic Transport in Dynamical Systems. Springer-Verlag, New York, NY.Google Scholar
Wu, J. Z., Gu, J. W., and Wu, J. M. (1987). Steady three-dimensional fluid particle separation from arbitrary smooth surface and formation of free vortex layers. AIAA, Paper 87-2348.Google Scholar
Wu, J. Z., Tramel, R. W., Zhu, F. L., and Yin, X. Y. (2000). A vorticity dynamics theory of three-dimensional flow separation. Phys. Fluids, 12, 19321954.Google Scholar
Wu, J. Z., Ma, H. Y., and Zhou, M. D. (2005). Vorticity and Vortex Dynamics. Springer, New York, NY.Google Scholar
Yagasaki, K. (2008). Invariant manifolds and control of hyperbolic trajectories on infinite- or finite-time intervals. Dyn. Sys., 23, 309331.Google Scholar
Yamada, H., and Matsui, T. (1978). Preliminary study of mutual slip-through of a pair of vortices. Phys. Fluids, 21(2), 292294.Google Scholar
Yang, H., Weisberg, R. H., Niiler, P. P., Sturges, W., and Johnson, W. (1999). Lagrangian circulation and forbidden zone on the West Florida Shelf. Continental Shelf Research, 19, 12211245.Google Scholar
Yang, K., Wu, S., Zhang, H., et al. (2021). Lagrangian-averaged vorticity deviation of spiraling blood flow in the heart during isovolumic contraction and ejection phases. Med. Biol. Eng. Comput. 59, 14171430.Google Scholar
Yates, L. A., and Chapman, G. T. (1992). Streamlines, vorticity lines, and vortices around three-dimensional bodies. AIAA J., 30, 18191826.Google Scholar
Yuster, T., and Hackborn, W. W. (1997). On invariant manifolds attached to oscillating boundaries in Stokes flows. Chaos, 7(4), 769776.Google Scholar
Zhang, W., Wolfe, C. L. P., and Abernathey, R. (2020). Role of surface-layer coherent eddies in potential vorticity transport in quasigeostrophic turbulence driven by eastward shear. Fluids, 5(1), 2.Google Scholar
Zhang, X., Hadwiger, M., Theussl, T., and Rautek, P. (2022). Interactive exploration of physically-observable objective vortices in unsteady 2D flow. IEEE Trans. Vis. Comp. Graph., 28(1), 281290.Google Scholar
Zhang, Z., Wang, W., and Qiu, B. (2014). Oceanic mass transport by mesoscale eddies. Science, 345(6194), 322324.Google Scholar
Zhong, Y., Bracco, A., and Villareal, T. A. (2012). Pattern formation at the ocean surface: Sargassum distribution and the role of the eddy field. Limnol. Oceanog., 2(1), 1227.Google Scholar
Zhou, J., Adrian, R. J., Balachandar, S., and Kendall, T. M. (1999). Mechanisms for generating coherent packets of hairpin vortices in channel flow. J. Fluid Mech., 387, 353396.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • George Haller, ETH Zurich
  • Book: Transport Barriers and Coherent Structures in Flow Data
  • Online publication: 20 February 2023
  • Chapter DOI: https://doi.org/10.1017/9781009225199.012
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • George Haller, ETH Zurich
  • Book: Transport Barriers and Coherent Structures in Flow Data
  • Online publication: 20 February 2023
  • Chapter DOI: https://doi.org/10.1017/9781009225199.012
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • George Haller, ETH Zurich
  • Book: Transport Barriers and Coherent Structures in Flow Data
  • Online publication: 20 February 2023
  • Chapter DOI: https://doi.org/10.1017/9781009225199.012
Available formats
×