1. Introduction
Existing evidence based on observational estimates and numerical simulations, however, suggests that the eddy-induced transport is spatially inhomogeneous (e.g., LaCasce and Bower 2000) and anisotropic, that is, it has a preferred direction (e.g., Freeland et al. 1975; Spall et al. 1993; LaCasce 2000). The along-isopycnal eddy diffusivity can be described by a location-dependent two-dimensional tensor, and the preferred direction can be determined by diagonalizing this tensor. The latter approach was taken by Rypina et al. (2012), who analyzed trajectories of both synthetic Lagrangian particles (diagnosed from the altimetric data) and the actual surface drifters in the North Atlantic. The results demonstrate that the preferred transport direction varies across the region, the transport anisotropy is caused primarily by geostrophic rather than nongeostrophic currents (see also Sallee et al. 2008), and the spreading of Lagrangian particles can be faster or slower than diffusive, that is, “superdiffusive” or “subdiffusive,” respectively (see also Berloff et al. 2002; Veneziani et al. 2005; Kamenkovich et al. 2009).
The origins of this complexity remain largely unclear, and several mechanisms have been proposed. The mean advection can significantly modulate the eddy-induced transport. In particular, the meridional diffusivity is enhanced at steering levels (Green 1970; Killworth 1997) and is suppressed by zonal propagation of eddies relative to the mean zonal flow (Ferrari and Nikurashin 2010); meridional shear in zonal currents can cause shear dispersion (e.g., Taylor 1953; Young et al. 1982; Smith 2005); and cross-jet transport barriers exist on strong currents such as the Gulf Stream and its extension (Samelson 1992; Rypina et al. 2011) and alternating multiple jets (Haynes et al. 2007; Berloff et al. 2009). In addition, powerful mean currents, such as those within the western boundary regions and the upper-ocean Antarctic Circumpolar Current, can dwarf the along-stream eddy-induced transport.
In many parts of the ocean, however, mean currents are weak relative to eddies, and the along-stream diffusivity is as important for tracer distribution as the mean advection. In these regions, the anisotropy cannot be explained by the effects of the mean advection alone (Kamenkovich et al. 2009; Rypina et al. 2012). On the other hand, the eddy velocity variance tends to be isotropic (Rypina et al. 2012) and cannot explain anisotropy in K using (1) either. Kamenkovich et al. (2009) hypothesize that the dominance of the zonal eddy diffusivity can be caused by zonally elongated eddies such as those observed in altimetry-based observational datasets (Huang et al. 2007), and this hypothesis is further examined in this study. This manuscript investigates the influence of zonally elongated transient patterns on the particle spreading, describes spectral and transport properties of these transients in idealized numerical simulations (sections 2 and 3) and altimetry-based velocity estimates (section 4), and discusses the importance of transient motions in idealized tracer distribution in the model context in section 5.
2. Numerical model and simulated flow
The dynamical model is adapted from Karabasov et al. (2009) and only a very brief description of it is given here. This model employs an advanced advection scheme Compact Accurately Boundary-Adjusting High-Resolution Technique (CABARET) which allows achieving highly effective spatial resolution, meaning that numerical convergence is found at much coarser spatial resolution than in the case of traditional advection schemes. An equally important and attractive property of this formulation is its numerical stability in the presence of small dissipation, which allows simulations with very high, and most realistic, Reynolds numbers (Re).
The simulated flow and its spectrum
The simulated flow consists of subtropical and subpolar gyres, separated by a well-pronounced western boundary current and its eastward jet extension (EJE hereinafter; Fig. 1a). The entire domain is filled with mesoscale eddies, which are particularly strong in the vicinity of the EJE (Fig. 1b). The magnitudes of motions decrease with depth. The spatial structure of the PV is qualitatively similar to the streamfunction (Figs. 1c,d). This similarity is explained by the dominance of the stretching terms [last terms in (3)], which is because the dominant length scales in the solutions are several Rossby deformation radii (see the following discussion of Fig. 2).
Circulation in the top layer of the numerical simulations. (a) Time-mean (over 50 yr) streamfunction and (b) instantaneous minus the time-mean streamfunction (eddies) (m2 s−1). (c) Time-mean (over 50 yr) PV; (b) instantaneous minus the time-mean PV (eddies) (s−1).
Citation: Journal of Physical Oceanography 45, 3; 10.1175/JPO-D-14-0164.1
Spatial structure of the circulation in the numerical simulation. (a) Wavenumber k–l spectrum of velocity in the top layer, time averaged over 50 yr; absolute values of wavenumbers (k, l) are nondimensionalized by Rd1; the spectrum is nondimensionalized by the total kinetic energy (multiplied by 0.005). (b) The spectral power as a function of the angle between the wavenumber (k, l), summed over the interval K = [1/20Rd1 1/10Rd1] and divided by its maximum value. Note the presence of the anisotropic peak at small k in (a) and at angle ≈ 85° and in (b) corresponding to the zonal transients. (c) Zonal transients, isolated by the low-pass filtering (using the sine transform) of the instantaneous velocity streamfunction in the top layer (m2 s−1). (d) Zonally and time-averaged kinetic energy (weighted by the total kinetic energy) of the zonal transients in the three vertical layers.
Citation: Journal of Physical Oceanography 45, 3; 10.1175/JPO-D-14-0164.1
The spatial structure of the eddy field is illustrated by the two-dimensional wavenumber (k–l) spectrum of the velocity—the sum of u and υ velocity spectra (Fig. 2a). The 2800 spectra of instantaneous velocities are computed at 7-day intervals and then averaged in time; note that these instantaneous k–l spectra do not contain information on time dependence and eddy propagation. Most of the spectral power is contained in the circular band corresponding to the total wavenumbers
To study the meridional structure and propagation properties of zonal transients, we isolate them by spatial filtering of the velocity streamfunction; the filtering is done in the zonal direction only, and the cutoff wavelength is 30Rd1. As the rest of the flow, the zonal transients have maximum amplitudes in the EJE vicinity (Figs. 2c,d). With depth, the distribution of the kinetic energy of the zonal transients becomes more uniform in the meridional direction and the relative importance of zonal transients in the regions north and south of EJE is the largest in layer 3. We will later see that this deep region also corresponds to the largest transport anisotropy.
Zonal transients propagate westward at a speed of approximately 0.035–0.05 m s−1 north and south of EJE, as estimated from the Hovmöller diagrams. The phase speed of zonal transients is noticeably smaller than the phase speed of the barotropic Rossby wave with the same wavenumbers and in the motionless medium (0.09 m s−1) but are larger than the phase speed of the first baroclinic Rossby wave (0.02 m s−1). This discrepancy is likely to be explained by the effects of the mean advection (Berloff and Kamenkovich 2013a), but the analysis of the normal modes of the double-gyre flow is beyond the scope of this study.
3. Lagrangian analysis
The effects of the mean flow on the eddy-induced diffusivity is accounted for by the full trajectory-following (FTF) method (Berloff et al. 2002; Rypina et al. 2012), which was shown by the latter study to account for such known effects of the mean flow on the eddy diffusivity as the cross-jet suppression of eddy-induced particle spreading and material transport barriers. The method calculates particle dispersion only due to the time-dependent (eddy) part of the flow but along the particle trajectories in the full (eddy plus mean) flow. This effectively captures the effects of the mean advection on the eddy-driven dispersion because the Lagrangian quantities in (1) are determined by particle location. Note that the more straightforward analysis of particles in the full flow cannot serve this purpose because of the particle dispersion by the mean flow itself.
Neutrally buoyant Lagrangian particles are released in 50 consecutive 400-day segments, starting with 130 000 particles in each layer. To examine the spatial distribution of the anisotropic spreading rates, this area is divided into 106 km by 106 km subregions, and the particles are divided into the corresponding groups, according to their initial positions. Particle spreading rates are computed for each subregion over the 400-day time interval. Typically, most particles in each group leave the subregion boundaries before they reach the diffusive regime, and these nonlocal effects must be accounted for. To do this for each group, we define a mean “particle cloud” by its center of mass, using (X, Y), and by its size, using the average zonal/meridional displacements. If several particle clouds overlap at a given point, the dispersion at this point is estimated by the ensemble average of the corresponding individual cloud dispersions. Particle clouds that touch solid boundaries are discarded.
a. Dispersion regimes
The long time asymptotic behavior of (8) is traditionally used to characterize different dispersion regimes (e.g., LaCasce 2008). In particular, the diffusive regime corresponds to the linear increase of the dispersion with time, achieved after sufficient time has passed, and (8) then provides an estimate for the eddy diffusivities. Deviations from the diffusion are quantified here by fitting tα+1 to
The map of parameter α shows that over the 400 days used to estimate the diffusivity in this study, the dispersion is not exactly diffusive in most of the domain (Fig. 3). In particular,
Spreading regimes in the control simulation. Parameters (left)
Citation: Journal of Physical Oceanography 45, 3; 10.1175/JPO-D-14-0164.1
The importance of the mean advection can be estimated here by comparing the dispersion regimes in the control FTF simulation and a more traditional “eddy-only” run, in which the effects of the mean advection are neglected, and the particles only feel (i.e., are advected by) the eddying component of the flow. The eddy-only simulation exhibits more diffusive spreading along the main axis and the basin-averaged magnitude of
b. Anisotropic dispersion
The eddy diffusivity is strongly anisotropic, with Kξ exceeding Kη everywhere in the domain (Fig. 4a; Table 1). The largest diffusivities are found between the gyres in the EJE-dominated part of the domain, where the eddy kinetic energy is also the highest. In the top layer, the anisotropy parameter
Anisotropic spreading rates in the control simulation. (left) Spreading ellipses (see text) are superimposed here on the anisotropy parameter aaniso (shaded); every ninth ellipse is shown for presentation purposes. Also shown is the time-mean streamfunction. (right) Zonally averaged aaniso in the three vertical layers.
Citation: Journal of Physical Oceanography 45, 3; 10.1175/JPO-D-14-0164.1
Anisotropy coefficient aaniso in four simulations, area averaged within three vertical layers.
The correlation time scales in the major and minor directions
Correlation time scale
c. Causes of the dispersion anisotropy: Mean advection and zonal transients
We first examine the role of the mean advection by comparing the control simulation to the eddy-only (EO) experiment (which was described in section 3a). In the EO simulations, the anisotropy parameter aaniso decreases in the top layer, with the largest changes in the EJE vicinity (not shown). However, aaniso remains larger than 2.0 in most of the domain and is larger than 5 in the northern and southern parts of the domain; the area-averaged value is 4.0. This demonstrates that, even in the absence of mean advection, the eddies cause anisotropic particle spreading. Because of the weakness of the mean advection in the deep layers, the differences between the standard and EO runs are only noticeable in the EJE vicinity. Interestingly,
We next estimate the importance of zonal transients by analyzing a “zonal transient–dominated” sensitivity experiment (Fig. 5a). In this run, we low-pass Fourier filter the velocity streamfunction in the zonal direction with Lfilter = 30Rd1 (simulation LPx30Rd). For this purpose, the flow is decomposed into the Fourier series,1 all Fourier coefficients corresponding to scales shorter than Lfilter are set to zero, and the inverse transform is applied. This simulation employs the FTF technique, so the full trajectories of particles are the same as in the control simulation.
Sensitivity runs with the Fourier-filtered flows and the importance of zonal transients on the anisotropic spreading rates. (a) Zonal transient–dominated LPx30Rd simulation; (b) isotropic eddy–dominated HPx30Rd simulation. Spreading ellipses (see text) are superimposed here on the anisotropy parameter aaniso (shaded); every ninth ellipse is shown for presentation purposes. Also shown is the time-mean streamfunction.
Citation: Journal of Physical Oceanography 45, 3; 10.1175/JPO-D-14-0164.1
The spreading rates become strongly anisotropic with aaniso exceeding 10.0 in most of the domain; aaniso also becomes more spatially uniform. Both
We now reverse the sensitivity experiment and carry out a simulation in the “isotropic eddy–dominated” experiment (HPx30Rd), where zonal transients are removed from the velocity streamfunction using the high-pass Fourier filter with Lfilter = 30Rd1. Several differences with the control simulation are notable. First, the spreading becomes more isotropic with the area-averaged aaniso ≈ 2.0 in all layers. This is despite the fact that
Correlation time scale
4. Anisotropic transport and its causes in altimetry-based estimates
The model-based results in section 3 strongly indicate that the anisotropy of the eddy-induced material transport and the predominantly zonal direction of preferred particle spreading are largely controlled by zonal transients. We now test these conclusions using a 17-yr-long record (from 1992 to 2009) of the geostrophic velocities inferred from AVISO sea surface height altimetric measurements. We focus here on the subtropical North Atlantic from 20° to 50°N and from 70° to 20°W; the data and methods are the same as in Rypina et al. (2012). Similar to the model-based k–l velocity spectrum shown in Fig. 2, the spectrum of geostrophic velocities (Fig. 6) contains a noticeable peak in its zonal transient portion, where zonal scales exceed 1000 km. Unlike the model results, however, the isotropic part of the spectrum contains multiple peaks.
Spatial (k–l) velocity spectrum of the geostrophic velocity (sum of the u and υ velocity spectra, where u and υ are in km day−1) inferred from the AVISO satellite altimetry, time averaged over the period from 1992 to 2009.
Citation: Journal of Physical Oceanography 45, 3; 10.1175/JPO-D-14-0164.1
We now investigate the influence of this zonal transient spectral peak on the eddy-induced diffusivity by comparing particle spreading in simulations with the unfiltered eddies (control run) to simulations with the low-pass filtered (zonal transient dominated) and high-pass filtered (isotropic eddy dominated) eddy fields. As before, the diffusivities are quantified using the FTF approach and are visualized using the diffusivity ellipses (Fig. 7). In comparison to the control run (green ellipses), in the zonal transient–dominated simulations (blue ellipses) both the zonal and meridional components of diffusivity become smaller, but the meridional component decreases significantly more than the zonal component. As a result, the ellipses become nearly zonal throughout most of the domain, and the anisotropy coefficient increases from 5.4 in the control to 7.9 in the zonal transient–dominated run. If, in the opposite, zonal transients are removed in the isotropic eddy–dominated flow, the zonal component of diffusivity decreases more than the meridional, and the ellipses become less anisotropic with the domain-averaged anisotropy coefficient of only 2.5 (Fig. 7, bottom). All of these results are in agreement with the model-based results of section 3.
Anisotropic transport and its causes in altimetry-based estimates of North Atlantic circulation. (top) Diffusivity ellipses in the three simulations: full unfiltered flow (green), low-pass filtered zonal transient–dominated flow (blue), and high-pass filtered flow (red). Anisotropy parameter aaniso in three simulations: (bottom left) full unfiltered flow, (bottom middle) low-pass filtered flow, and (bottom right) high-pass filtered flow.
Citation: Journal of Physical Oceanography 45, 3; 10.1175/JPO-D-14-0164.1
5. Tracer distribution in the numerical model



Importance of eddies in the idealized tracer distributions. Tracer concentrations are shown at day 100 for two simulations with F = 0 and (a) full flow and (b) mean advection only. Central patch release is not shown because of its strong deformation by the EJE. Time-mean streamlines are shown by the black contours.
Citation: Journal of Physical Oceanography 45, 3; 10.1175/JPO-D-14-0164.1
In the control simulation, the tracer is advected by the full flow (mean and eddy) and
Distribution of an idealized tracer in numerical simulations. Tracer patches from the three different releases (northern patch, central patch, and southern patch) are overlapped and shown (in the top layer at day 200) for the initial distribution and sensitivity experiments.
Citation: Journal of Physical Oceanography 45, 3; 10.1175/JPO-D-14-0164.1
Weighted mean square error in the tracer concentration (see text) at day 200 shown for the three tracer patches and in the three vertical layers.
The eddy velocities are removed and replaced with
What is the relative importance of zonal transients and isotropic eddies? To answer this question, we carried out the zonal transient–dominated simulation with LPx30Rd velocities and isotropic eddy–dominated simulation with HPx30Rd velocities. Both simulations have
The effects of zonal transients are further studied in the zonal transient–dominated run. Simulated tracer distributions are surprisingly close to the simulation with
6. Discussion and conclusions
This study examines the anisotropic transport properties of the eddying North Atlantic flow, using an idealized model of the double-gyre oceanic circulation and altimetry-derived velocities. In this study, we decompose the flow into three main components: time-mean advection, large-scale zonal transients, and the remainder of the eddy field. The material transport by the time-dependent flow (quantified by the eddy diffusivity tensor) varies geographically and is anisotropic, that is, it has a well-defined direction of the maximum transport. These properties are primarily explained by the action of transient motions, rather than the effects of the time-mean advection. In particular, zonal transients correspond to the primarily zonal material transport and explain the largest part of anisotropy in diffusivities for both numerically simulated and altimetry-based velocity fields.
Zonal transients are defined using the spatial velocity spectrum, which, in the upper ocean, shows a peak at the basinwide zonal scale and a nearly meridional wavevector. Because of these spectral properties, Lagrangian velocities in zonal transient–dominated flows are predominantly zonal and have persistent correlations in time. This makes zonal transients a particularly effective vehicle for the anisotropic material transport, despite the fact that the amount of energy contained in the zonal transient portion of the spectrum is relatively small. Anisotropy in transport is due primarily to the difference in the correlation time scales, rather than anisotropy of the velocity covariance matrix. Our definition of these transients is based solely on their zonal scales and they are, strictly speaking, spectral Fourier modes in the zonal direction. The dynamical interpretation of these transients and their origins remains to be established. In particular, it is possible that zonal transients are normal modes and exist because of the linear dynamics through their interactions with the mean flow (Berloff and Kamenkovich 2013a,b). Alternatively, the energy at the zonal transient part of the spectrum can exist because of the nonlinear energy transfer due to interactions among transient eddies (Arbic et al. 2014). Investigation of the dynamics of zonal transients is left for future studies.
Anisotropy in transport is quantified here using a diagonalized diffusivity tensor, although the transport properties are almost never perfectly diffusive. This nondiffusive behavior, combined with spatial inhomogeneity and anisotropy, makes the parameterization of eddy-induced transport challenging. This is demonstrated by biases in idealized tracer distributions in simulations, in which the eddy-induced transport is parameterized using Lagrangian diffusivity estimates. Since such estimates are not globally available below the surface, finding an effective parameterization for the entire eddying flow may be even more difficult than our study implies. Our results suggest, however, that this task becomes easier in simulations with explicit zonal transients since these flow components are associated with a large part of the complexity in the transport, such as spatial variability in the decorrelation scales and anisotropy. Zonal transients are large enough to be resolved by most numerical simulations even at relatively coarse spatial resolution, but such non-mesoscale-resolving simulations may lack the dynamics necessary to simulate zonal transients. The importance of large-scale transients and the utility of the Lagrangian estimates of eddy diffusivity need to be further studied for more realistic, climatically relevant tracers. This can be done using simulations with and without eddy advection (as in Booth and Kamenkovich 2008) and will be a subject of a future study.
Acknowledgments
We thank two anonymous reviewers for their helpful suggestions on improving this manuscript. IK would like to acknowledge support through the NSF Grant OCE-1154923. IR was supported by the NSF OCE-1154641 and NASA Grant NNX14AH29G.
REFERENCES
Arbic, B., M. Mueller, J. Richman, J. Shriver, A. Morten, R. Scott, G. Serazin, and T. Penduff, 2014: Geostrophic turbulence in the frequency–wavenumber domain: Eddy-driven low-frequency variability. J. Phys. Oceanogr., 44, 2050–2069, doi:10.1175/JPO-D-13-054.1.
Berloff, P., and I. Kamenkovich, 2013a: On spectral analysis of mesoscale eddies. Part I: Linear analysis. J. Phys. Oceanogr., 43, 2505–2527, doi:10.1175/JPO-D-12-0232.1.
Berloff, P., and I. Kamenkovich, 2013b: On spectral analysis of mesoscale eddies. Part II: Nonlinear analysis. J. Phys. Oceanogr., 43, 2528–2544, doi:10.1175/JPO-D-12-0233.1.
Berloff, P., J. C. McWilliams, and A. Bracco, 2002: Material transport in oceanic gyres. Part I: Phenomenology. J. Phys. Oceanogr., 32, 764–796, doi:10.1175/1520-0485(2002)032<0764:MTIOGP>2.0.CO;2.
Berloff, P., I. Kamenkovich, and J. Pedlosky, 2009: A mechanism for the formation of multiple zonal jets in the oceans. J. Fluid Mech., 628, 395–425, doi:10.1017/S0022112009006375.
Booth, J., and I. Kamenkovich, 2008: Isolating the role of mesoscale eddies in mixing of a passive tracer in an eddy resolving model. J. Geophys. Res., 113, C05021, doi:10.1029/2007JC004510.
Ferrari, R., and M. Nikurashin, 2010: Suppression of eddy diffusivity across jets in the Southern Ocean. J. Phys. Oceanogr., 40, 1501–1519, doi:10.1175/2010JPO4278.1.
Freeland, H. J., P. B. Rhines, and T. Rossby, 1975: Statistical observations of the trajectories of neutrally buoyant floats in the North Atlantic. J. Mar. Res., 33, 383–404.
Green, J. S. A., 1970: Transfer properties of the large-scale eddies and the general circulation of the atmosphere. Quart. J. Roy. Meteor. Soc., 96, 157–185, doi:10.1002/qj.49709640802.
Haynes, P. H., D. A. Poet, and E. F. Shuckburgh, 2007: Transport and mixing in kinematic and dynamically consistent flows. J. Atmos. Sci., 64, 3640–3651, doi:10.1175/JAS4030.1.
Henning, C. C., and G. Vallis, 2004: The effects of mesoscale eddies on the main subtropical thermocline. J. Phys. Oceanogr., 34, 2428–2443, doi:10.1175/JPO2639.1.
Huang, H.-P., A. Kaplan, E. Curchitser, , and N. Maximenko, 2007: The degree of anisotropy for mid-ocean currents from satellite observations and an eddy-permitting model simulation. J. Geophys. Res., 112, C09005, doi:10.1029/2007JC004105.
Kamenkovich, I., P. Berloff, and J. Pedlosky, 2009: Anisotropic material transport by eddies and eddy-driven currents in a model of the North Atlantic. J. Phys. Oceanogr., 39, 3162–3175, doi:10.1175/2009JPO4239.1.
Karabasov, S. A., P. S. Berloff, and V. M. Goloviznin, 2009: CABARET in the ocean gyres. Ocean Modell., 30, 155–168, doi:10.1016/j.ocemod.2009.06.009.
Killworth, P., 1997: On the parameterization of the eddy transfer. Part I. Theory. J. Mar. Res.,55, 1171–1197, doi:10.1357/0022240973224102.
LaCasce, J. H., 2000: Floats and f/H. J. Mar. Res., 58, 61–95, doi:10.1357/002224000321511205.
LaCasce, J. H., 2008: Lagrangian statistics from oceanic and atmospheric observations. Transport and Mixing in Geophysical Flows, J. B. Weiss and A. Provezale, Eds., Springer, 165–228.
LaCasce, J. H., and A. Bower, 2000: Relative dispersion in the subsurface North Atlantic. J. Mar. Res., 58, 863–894, doi:10.1357/002224000763485737.
Prandtl, L., 1925: Bericht über Untersuchungen zur ausgebildeten Turbulenz. Z. Angew. Math. Mech., 5, 136–139.
Rypina, I. I., L. J. Pratt, and M. S. Lozier, 2011: Near-surface transport pathways in the North Atlantic Ocean: Looking for throughput from the subtropical to subpolar gyre. J. Phys. Oceanogr., 41, 911–925, doi:10.1175/2011JPO4498.1.
Rypina, I. I., I. Kamenkovich, P. Berloff, and L. Pratt, 2012: Eddy-induced particle dispersion in the near-surface North Atlantic. J. Phys. Oceanogr., 42, 2206–2228, doi:10.1175/JPO-D-11-0191.1.
Sallee, J.-B., K. Speer, R. Morrow, and R. Lumpkin, 2008: An estimate of Lagrangian eddy statistics and diffusion in the mixed layer of the Southern Ocean. J. Mar. Res., 66, 441–463, doi:10.1357/002224008787157458.
Samelson, R., 1992: Fluid exchange across a meandering jet. J. Phys. Oceanogr., 22, 431–444, doi:10.1175/1520-0485(1992)022<0431:FEAAMJ>2.0.CO;2.
Smith, K. S., 2005: Tracer transport along and across coherent jets in two-dimensional turbulent flow. J. Fluid Mech., 544, 133–142, doi:10.1017/S0022112005006750.
Spall, M. A., P. L. Richardson, and J. Price, 1993: Advection and eddy mixing in the Mediterranean salt tongue. J. Mar. Res., 51, 797–818, doi:10.1357/0022240933223882.
Taylor, G. I., 1921: Diffusion by continuous movements. Proc. London Math. Soc., 20, 196–211, doi:10.1112/plms/s2-20.1.196.
Taylor, G. I., 1953: Dispersion of soluble matter in solvent flowing slowly through a tube. Proc. Roy. Soc. London, A219, 186–203, doi:10.1098/rspa.1953.0139.
Vallis, G. K., 2006: Atmospheric and Oceanic Fluid Dynamics. Cambridge University Press, 745 pp.
Veneziani, M., A. Griffa, A. M. Reynolds, Z. D. Garraffo, and E. P. Chassignet, 2005: Parameterizations of Lagrangian spin statistics and particle dispersion in the presence of coherent vortices. J. Mar. Res., 63, 1057–1083, doi:10.1357/002224005775247571.
Young, W. R., P. B. Rhines, , and C. J. R. Garrett, 1982: Shear-flow dispersion, internal waves and horizontal mixing in the oceans. J. Phys. Oceanogr., 12, 515–527, doi:10.1175/1520-0485(1982)012<0515:SFDIWA>2.0.CO;2.
Note that although the flow satisfies the no-normal flow and no-slip boundary conditions, the streamfunction is not periodic in the strict sense. Nevertheless, the results with the Fourier transform and with and without window tapering and the use of the sine transform lead to very similar results.