1870 JOURNAL OF PHYSICAL VOLUME 32

A Laboratory Model of Exchange and Mixing between Western Boundary Layers and Subbasin Recirculation Gyres*

HEATHER E. DEESE,LARRY J. PRATT, AND KARL R. HELFRICH Woods Hole Oceanographic Institution, Woods Hole, Massachusetts

(Manuscript received 25 January 2001, in ®nal form 8 November 2001)

ABSTRACT Chaotic advection is suggested as a possible mechanism for ¯uid exchange and mixing among a western boundary current and subbasin recirculation gyres. Applications include the North Atlantic Deep Western Bound- ary Current and its adjacent mesoscale recirculation gyres. Visualization and quanti®cation of certain aspects of chaotic advection in a laboratory analog are described. Depending on the strength of the forcing, recirculating ¯uid offshore of the western boundary layer may be contained in a single gyre (not favorable for chaotic advection) or twin gyre with a ``®gure-eight'' geometry (favorable for chaotic advection). When time dependence is imposed on these steady ¯ows by varying the forcing periodically, the resulting ¯uid exchange, stirring, and mixing is most dramatic in the case of the twin gyre. A template for these processes can be formed by highlighting certain material contours (invariant manifolds) using dye and other techniques. These objects can be used to identify blobs of ¯uid (turnstile lobes) that are carried into and out of the gyres. The associated transports and ¯ushing times can be estimated. The preferential stirring and mixing in the twin-gyre case is quanti®ed by calculating the effective diffusivity of the ¯ow ®eld based on snapshots of the dye ®elds at longer times. The experiment suggests how tracers in a western boundary current might be transported into and out of neighboring recirculations and where regions of strong zonal or meridional transport might occur.

1. Introduction are on the order of 1 cm sϪ1, while mean observed velocities are on the order of 5±10 cm sϪ1 (e.g., Watts Western boundary currents exchange mass, tracers, 1991).] The discrepancy could be due to the fact that and dynamical properties with subbasin-scale recircu- DWBC water is temporarily captured in recirculations lations. Observations from the deep North Atlantic have during its southward transit (Bower and Hunt 2000a) or revealed the presence of numerous mesoscale recircu- that the water adjacent to the DWBC is not tracer-free lations along the offshore ¯ank of the Deep Western (Pickart et al. 1989; Rhein 1994). There are also many Boundary Current (DWBC). Examples include meso- well-known surface gyres, such as the Great Whirl, the scale gyres in the Labrador Sea (Lavender et al. 2000, Alboran Gyre, and the inertial recircula- hereafter LRO) near the Gulf Stream undercrossing tions, that coexist and interact with surface boundary (Hogg 1992; Bower and Hunt 2000a,b), to the northeast currents. of the Bahamas (Lee et al. 1996; Leaman and Vertes Tracers, ¯oats, geostrophic calculations, and current 1996; Johns et al. 1997), and in the Guiana Basin (Mc- meter velocities have been used to construct rough maps Cartney 1992; Friedrichs and Hall 1993). Schmitz and of the features mentioned above. An example from the McCartney (1993) summarize some of these features North Atlantic DWBC appears in Fig. 1 (after Johns et along with deep basin-scale gyres. Fine (1995) has al. 1997), which shows the time-average streamline ®eld shown that tracer-derived ages for the DWBC water are for the deep ¯ow in the vicinity of Abaco Island. The an order of magnitude older than the direct transit time ®gure suggests twin gyres just offshore of the main from the source regions based on directly observed ve- southward ¯ow. (Note, however, that this feature is locities in the core of the ¯ow. [Tracer derived velocities based largely on geostrophic calculations subject to ref- erence velocity uncertainties.) An annual DWBC trans- * Woods Hole Oceanographic Institution Contribution Number port cycle and 30±150-day onshore/offshore excursions 10450. of the DWBC have been detected (Lee et al. 1996), showing that this region is quite time dependent. A sub- stantially more complicated group of deep recirculations Corresponding author address: Larry Pratt, Woods Hole Ocean- ographic Institution, Mail Stop 21, 360 Woods Hole Rd., Woods Hole, in the Labrador Sea has been inferred by LRO using MA 02543. deep and intermediate ¯oat data. The dynamical factors E-mail: [email protected] leading to the formation of deep gyres are unknown,

᭧ 2002 American Meteorological Society JUNE 2002 DEESE ET AL. 1871 but a number of in¯uences including topographic steer- ing, inertial effects, and interactions with shallower lay- ers are possible. Thompson (1995) and references there- in discuss some of these processes within the context of numerical models. The focus of this work is not on the dynamics of recirculations but rather on a mechanism for exchange and mixing among recirculations and boundary currents. Such exchange is sometimes modeled as a diffusion process (Pickart and Hogg 1989) in which tracers cross the closed boundaries of steady recirculations. This view of tracer spreading is consistent with conventional ideas regarding classical turbulence, in which small-scale ed- dies give rise to coarse grain diffusion of passive tracers. However the observed geometry of deep recirculations and the presence of time dependence suggest that cha- otic advection could also be a factor. This process, which is thought to account for the transport of chemical spe- cies in the stratospheric polar vortex (Haynes 1999), involves advective mass exchange between regions of fundamentally different motion due to time dependence. FIG. 1. Approximate streamlines of the deep ¯ow below potential temperature 6ЊC, or about 1000-m depth. The contour interval is 2 The mass that is exchanged typically undergoes rapid Sverdrup (Sv ϵ 106 m3 s Ϫ1). Possible locations of hyperbolic tra- ®lamentation, a process that leads to intensi®cation of jectories are labeled ``H''. Adopted with permission from a ®gure of property gradients and is therefore a catalyst for irre- Johns et al. (1997). versible mixing through diffusion or subgrid-scale tur- bulence. Chaotic advection is generally associated with the occurrence of strong hyperbolicity (stretching and effective diffusivity (Nakamura 1996) measured from contraction about a certain parcel trajectory). The ten- dye distributions. The laboratory model is barotropic dency of closely spaced pairs of parcels to separate from and the dynamics therefore differ somewhat from those each other exponentially in time as they pass such a governing abyssal ¯ows. However, chaotic advection is trajectory is a fundamental characteristic of chaotic mo- a generic mechanism depending largely on the under- tion. lying geometry of the ¯ow. Differences between stirring In a steady ¯ow, a hyperbolic trajectory is simply a patterns observed in the laboratory model may be help- saddle-type stagnation point. Such points occur in the ful in understanding exchanges between boundary cur- Abaco recirculation system, approximately at the lo- rents and subbasin recirculations in general and in sug- cations marked H in Fig. 1. An idealized sketch of a gesting strategies for observing exchange processes in similar recirculation pair (Fig. 2) indicates a single hy- the ®eld. A moderate volume of literature on chaotic perbolic point at the self-intersection of the ®gure-eight advection has developed over the past 15 yr, and some separatrix. Recirculations need not have such points, as of our ultimate conclusions could have been anticipated shown for the single gyre sketched in Fig. 2b. If a time- on the basis of this work. However, laboratory models dependent perturbation is imposed, chaotic advection of chaotic advection are quite rare and the particular might be expected to occur more readily and with great- application to western boundary layer recirculations is er effect in the twin-gyre case (or in other geometries novel. having saddle-type stagnation points) than in the single- Section 2 reviews some of the basic objects and ter- gyre case. minology used in connection with chaotic advection, The above remarks suggest that ¯uid exchange and including hyperbolic trajectories and their stable and mixing along the DWBC may depend upon the under- unstable manifolds. Such objects can be used to con- lying geometry of the adjacent recirculations. In partic- struct a template for exchange and mixing among dis- ular, more extensive and intense mixing would seem to tinct regions of a ¯ow ®eld. Section 3 describes the be favored by geometries containing strong hyperbol- laboratory apparatus, experiment, and dimensionless pa- icity. The purpose of our work is to verify this conjecture rameters. (Brie¯y, the ¯ow is set up in a sliced cylinder by visualizing and quantifying ¯uid stirring and mixing with a differential lid rotation. The steady-state ¯ow in a laboratory model of a western boundary layer with consists of a Sverdrup interior, a western boundary layer, neighboring single- or twin-recirculation gyres. In cases and a single- or twin-recirculation gyre. The geometry where ¯uid exchange occurs, we would like to deter- of the gyre is determined by the strength of inertia, mine the pathways by which ¯uid enters and leaves the ultimately controlled by the differential lid rotation. gyre(s). Visualization of the ¯ow is achieved using dye Time dependence is induced by periodically varying the and particle imaging, and the mixing is quanti®ed using rate of differential rotation.) Section 4 is concerned 1872 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

FIG. 2. (a) and (b) Show steady streamline patterns for hypothetical twin- and single-gyre recirculations near a western boundary current. (c) The stable and unstable manifolds of the central hyperbolic trajectory for the time-dependent twin gyre. Stable manifolds are solid and unstable manifolds are dashed. The letters A, B, etc. denote material lobes of ¯uid that are being transport out of the northern gyre. AЈ, BЈ, etc. denote lobes containing ¯uid that is being transported into the northern gyre. (d) The recirculation boundary (thick solid line) formed by joining sections of the (dashed) unstable and (thin solid) stable manifolds. This boundary evolves as a material contour from t ϭ 0tot ϭ T Ϫ at which time it is rede®ned (t ϭ T). mainly with visualization of the stirring and exchange blobs. Section 6 contains a summary and discussion of process using dye. The dye patterns show profound dif- rami®cations of the work. ferences between the single- and twin-gyre cases and suggest where and how, in the twin-gyre case, recir- 2. Hyperbolic trajectories, manifolds, and lobe culating ¯uid is exchanged and mixed with surrounding dynamics ¯uid. Section 5 quanti®es the mixing process in terms of effective diffusivity, a bulk measure of the enhance- In order to describe the transport and stirring pro- ment of diffusion by advective ®lamentation of dyed cesses associated with chaotic advection, it is helpful JUNE 2002 DEESE ET AL. 1873 to use certain Lagrangian objects including hyperbolic The exchange and stirring mechanism in the unsteady trajectories, stable and unstable manifolds, and turnstile ¯ow is revealed by considering a lobe of ¯uid (marked lobes. We use these material objects to construct a tem- A in Fig. 2c) trapped between the stable and unstable plate for exchange processes due to the time dependence manifolds. Since the manifolds are material curves the of the ¯ow and to measure the associated volume ¯uxes. ¯uid in such a lobe must remain trapped between the The following is a quanlitative review; more complete manifolds and must preserve area as the ¯ow evolves. and formal discussions can be found in Rom-Kedar and In general, the lobe will follow the direction of the Wiggins (1990) and Wiggins (1992). It is generally as- cyclonic swirling ¯ow around the recirculation and will, sumed that the ¯ow in question is 2D and horizontally after a certain time interval, lie at cross-hatched position nondivergent, or at least so to a ®rst approximation. B, then C, and so on. Since Fig. 2c shows the manifolds We have previously referred to Figs. 2a,b depicting at a particular time, the lobes at B, C, D, etc. do not steady twin- and single-gyre recirculations. Note that necessarily depict the result of the evolution of the lobe the saddle-type ®xed point in Fig. 2a has stable and at A. However, the reader may be able to imagine the unstable directions (as indicated by the arrows) corre- evolution as the lobe moves around the periphery of the sponding to the directions of contraction and stretching. recirculation from A to something like F, a process that The streamlines extending away from the saddle point has two profound results. First, the ¯uid at A, initially along these directions form the ®gure-eight separatrix inside of the recirculation, has clearly moved outside of that de®nes the boundaries of the recirculations. If these the recirculation at F. (At the same time, ¯uid outside hypothetical ¯ows are perturbed in a time-dependent moves inside by following something like the cross- manner, and if the perturbation is not too large or rapid, hatched lobes AЈ, BЈ, etc.) Second, the evolving lobe is then generalizations of the stagnation points will exist. stretched into a ®lament, such as the one at F. Fluid is The generalization of the saddle-type ®xed point in Fig. thereby stretched and folded as it is drawn into and out 2a is a distinguished hyperbolic trajectory (DHT), a ¯uid of the recirculation, a process that increases the spatial parcel path having stable and unstable directions (i.e., gradients of any materially conserved properties. The directions of strong contraction and stretching) for all exchange process thus acts as a catalyst for mixing by time. The location of this parcel and the orientation of sharpening gradients so that diffusion can act more rap- the stable and unstable directions will vary with time, idly. A similar mechanism can generally be expected to but the trajectory will remain in the vicinity of the steady act in connection with the southern gyre. saddle point. If the Eulerian time dependence is periodic with pe- If dye is continuously introduced at the (moving) lo- riod T, as is nearly the case in our laboratory experiment, cation of the DHT, it will leave the area along the two the shapes of the manifolds recur on period T. Lobe A unstable directions forming two material contours, the then evolves exactly into lobe B, then C, and so on. unstable manifolds. Another set of material contours Although this progression may not always be in se- consists of ¯uid that is attracted to the DHT and ap- quence (e.g., lobe A may map to C and then E, while proaches it along the stable directions asymptotically in B maps to D and then F), sequential mapping is ob- time. In numerical simulations, these stable manifolds served in the experiment. In this case, A, B, C, etc., can be found by introducing ``dye'' at the DHT and must have the same area. As we will demonstrate below, running time backward. Hypothetical stable and unsta- A and AЈ (and therefore BЈ, CЈ, etc.) must also have ble manifolds for a DHT resulting from the ®gure-eight equal areas. geometry are sketched in Fig. 2c. The solid and dashed The above discussion has tacitly assumed that the curves are the stable and unstable manifolds, respec- velocity ®eld is known over in®nite time and that the tively, at a particular time.1 Unlike Fig. 2a, where the hyperbolic trajectories persist over that time. If the time- stable and unstable manifolds correspond to the same dependent motions are violent enough, hyperbolicity material curves, the manifolds for the unsteady case are can become intermittent; a situation that apparently does distinct curves that often tangle about each other. The not arise in the laboratory experiment described below. tangle is evidence that chaotic Lagrangian motion exists A more immediate complication is that the velocity ®eld in the vicinity of the intersecting manifolds. A funda- is known only over ®nite time, as is the case in all mental characteristic of this motion is that nearby ¯uid laboratory, numerical, or ®eld experiments with non- parcels separate from each other exponentially in time, periodic time dependence. Finite time duration leads to implying rapid stirring of the ¯uid. Even in systems uncertainty in de®ning DHTs; a continuous group of where chaos cannot formally be de®ned, the presence nearby trajectories may all exhibit strong hyperbolicity of tangled manifolds still implies exponential separation (stretching and contraction) over the time record. In of ¯uid parcels over ®nite times. practice, one can usually select a particular hyperbolic trajectory from this group, approximate its stable and unstable manifolds, and perform the same analysis as 1 Normally, the term ``manifold'' refers to a surface in (x, y, t) space consisting of material that asymptotically approaches or departs before. The investigator is simply mapping out conve- from the DHT. The curves in Fig. 2c, often referred to as manifolds nient material curves that reveal particular Lagrangian themselves, are actually time slices through these surfaces. structures and transport in the ¯ow ®eld. The reader 1874 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

TABLE 1. Constant dimensional and nondimensional values for laboratory tank.

L ϭ 42.5 cm ␤ ϭ 3 (ms)Ϫ1 ␤ ϭ fs/H D ϭ 20 cm ␤* ϭ 0.3 ␤* ϭ ␤L/ f 1/3 s ϭ 0.15 ␦M ϭ 0.7 cm ␦M ϭ (␷ /␤) Ϫ1 ⍀ϭ2 rad s ␦S ϭ 0.5 cm ␦S ϭ r/␤ ϭ ␦E f /␤D Ϫ1 1/2 f ϭ 2⍀ϭ4 rad s ␦E ϭ 0.07 cm ␦E ϭ (␷ /⍀) 2 Ϫ1 1/2 1/2 ␷ ϭ 0.01 cm s E ϭ 0.0035 E ϭ ␦E/D

0 (shaded in Fig. 2d) has moved to lobe D at t ϭ T. Lobe C lies inside the boundary at t ϭ 0 but D lies outside the (rede®ned) boundary at t ϭ T. The result is a ``chaotic transport'' out of the gyre equal to the area of C divided by T. At the same time, material in lobe BЈ outside the boundary has moved to CЈ inside. Since the boundary recurs periodically, lobes BЈ and C (and therefore all other lobes) must have equal area. The swapping of ¯uid across the boundary of the recircu- lation has lead to the term turnstile lobes. Some inves- tigators prefer to think of the boundary as a ®xed curve de®ned by the union of the manifolds at t ϭ 0, a def- inition that does not alter the transport value. The move- ment of ¯uid across this boundary is continuous rather than discrete in time. However, use of a ®xed boundary is less meaningful if the time-dependence is aperiodic. FIG. 3. Sketch of rotating sliced cylinder with differentially The boundary of the time-dependent recirculation can rotating lid. be chosen in a variety of ways and this choice deter- mines where ¯uid enters and leaves the recirculation. If we had picked an intersection point other than I (Figs. who wishes to wade deeper into this subject is referred 2c or 2d) as the joining point, this would have led to a to Miller et al. (1997, 2002), Malhotra and Wiggins different de®nition of the interior of the recirculation (1998), Rogerson et al. (1999), and Ide et al. (2001). and different regions of entry and exit. However, some Stable and unstable manifolds can be used to de®ne choices lead to bizarre recirculations bearing little re- a formal Lagrangian recirculation boundary and to semblance to the corresponding mean or steady recir- quantify the volume ¯ux across that boundary. In Fig. culation. The usual practice is to choose such that the 2c, for example, a boundary might be de®ned (at t ϭ resemblance is strong. 0, say) as the curve that begins at the hyperbolic tra- jectory H and follows the dashed contour extending to 3. Lab apparatus and procedure the upper left. When the intersection point I with the solid curve is reached, one switches to the solid curve A barotropic western boundary current and adjacent and follows it back to the hyperbolic trajectory. The single or double recirculation gyres can be set up in a boundary at t ϭ 0 is thus the union of the dashed curve rotating ``sliced cylinder'': a cylindrical laboratory HI and solid curve IH. A subsegment of this boundary tank with a sloping bottom (Fig. 3). The relevant di- is shown as a darkened line in the t ϭ 0 frame of Fig. mensional and nondimensional parameters appear in 2d. As time progresses this closed material curve de- Tables 1 and 2. The discussion below and the appendix forms. If the time-dependence is periodic (with period describe the general procedures used but spares the T) however, the shapes of the manifolds themselves re- reader many of the technical details. The latter are cur at t ϭ T, 2T, etc. In this case the material intersection thoroughly covered in Deese (2001), available upon point I moves to another intersection, say J, as t pro- request to Dr. Pratt. gresses from 0 to T. Over the same time period the The tank is 42.5 cm in diameter, has a mean depth material boundary evolves into the darkened line in the of 20 cm, and is ®tted with a bottom of slope s ϭ 0.15. middle frame of Fig. 2d. At this point it is standard The actual depth therefore varies from 16.8 to 23.2 cm. practice to rede®ne the boundary to coincide with its The tank rotates at a ®xed rate ⍀ϭ2 rad sϪ1, leading position at t ϭ 0 (lower frame of Fig. 2d). The times to a topographic beta ␤ ϭ fs/H ϭ 3 (ms)Ϫ1. The cir- immediately before and after boundary rede®nition are culation is driven by a lid that rotates at a differential designated T Ϫ and T. As a result of this rede®nition, rate ⌬⍀ Ͻ 0, producing a uniform, anticyclonic surface some of the material inside the gyre now lies outside, stress curl. The surface stress results in the formation 1/2 and vice versa. Speci®cally, material in lobe C at t ϭ of an Ekman layer of thickness ␦E ϵ (␷/⍀) ϭ 0.07 JUNE 2002 DEESE ET AL. 1875

TABLE 2. Variable dimensional and nondimensional values for lab tank.

⌬⍀(1 sϪ1) 0.0076 0.0173 0.0480 0.2777a ⌬⍀ ⑀ ϭ (nd) 0.0038 0.0087 0.0240 0.0139 ⍀ ␦ ⌬⍀ ␦ ϭϭ1 8.1 (nd) 0.50 0.75 1.25 3.00 ␦M Ί ⍀ Ϫ1 wE ϭ ⌬⍀ ´ ␦E (cm s ) 0.00053 0.00120 0.00336 0.0194 Ϫ1 US ϭ wE/s (cm s ) 0.0035 0.008 0.022 0.129 b Ϫ1 Uwbc ϭ US ´ L/␦wbc (cm s ) 0.047 0.107 0.293 1.72 2 Ϫ5 Ϫ4 Ϫ4 Ϫ3 US/␤L 7.3 ϫ 10 1.7 ϫ 10 4.6 ϫ 10 2.7 ϫ 10 Re ϭ UL/␷ 14 32 88 516 Ϫ5 Ϫ5 Ϫ4 Ϫ4 Ro ϭ US/ fL 2.2 ϫ 10 5.0 ϫ 10 1.4 ϫ 10 8.1 ϫ 10 Ro ϭ ⌬⍀/2⍀ 0.0019 0.0044 0.012 0.070 ⑀/E 1/2 1.08 2.47 6.86 39.71 a The ¯ows for this setting were unsteady and were not used as a basis for any experimental runs. b Based on approximate observed value of western boundary layer thickness ␦wbc. cm (␷ being the kinematic of water) with a described in the appendix, this resolution is insuf®cient uniform vertical velocity wE ϭ⌬⍀␦E Ͻ 0 at its base. to perform a lobe analysis that adds signi®cantly to what Away from the dynamical western boundary this down- can be done with dye. Instead, PIV is used to make welling gives rise to a southward Sverdrup transport accurate velocity measurements over a limited area of with typical meridional velocity US ϭ wE/s. The Sver- the ¯ow ®eld in order to quantify the time dependence drup character of the regime is con®rmed by the small of the ¯ow. The third visualization technique involves 2 Ϫ4 values of Rossby number and US/␤L , both O(10 ). the introduction of dye into the ¯ow using a needle. Three types of boundary layers, the Stommel and Dye patterns are illuminated from below and recorded Munk frictional western boundary layers and the inertial from above using a video camera. The images are later boundary layer, all come into play at the dynamical digitized. The appendix contains further technical in- western boundary. The corresponding boundary layer formation regarding the particle and dye techniques. thicknesses are given by For cases of steady lid forcing, the (nearly) steady ¯ows2 observed agree with those found in laboratory 1/3 1/2 ␷␦EsfUexperiments with similar forcing and geometry (Ped- ␦ ϭ , ␦ ϭ , and ␦ ϭ . MS΂΃␤ D␤␤ I ΂ ΃ losky and Greenspan 1967; Beardsley 1969; Grif®ths and Kiss 1999). With ␦ less than about 1.1 the nearly As it turns out, ␦M (at 0.7 cm) and ␦S (at 0.5 cm) are steady ¯ow ®eld in western reaches of the tank consists quite close and constant in these experiments. We there- of a western boundary layer and a single recirculation 1/2 fore treat the ratio ␦ ϭ ␦I/␦M ϭ 8.1(⌬⍀/⍀) as the gyre. Within the range 1.1 Ͻ ␦ Ͻ 1.4, the recirculation primary indicator of nonlinearity near the western splits into twin gyres.3 Both steady regimes are revealed boundary layer. Our experiments are restricted to ␦ Յ by a sequence of streak images (Fig. 4) for the western 1.4, for which the ¯ow is steady, or very nearly so, in two-thirds of the tank. In this and other photos of the the presence of steady lid rotation. Time variation about circulation, the lower boundary is the western boundary the steady states is imposed by varying the lid rotation and north is to the left. Note that a shadow obscures a according to portion of the western boundary layer. The single-gyre ⌬⍀(T) ϭ⌬⍀[1 ϩ A sin(2␲t/T )]. geometry occurs for ␦ ϭ 1.00 (Fig. 4a). If ␦ is increased o osc osc past about 1.1 (Fig. 4b) the gyre splits forming the ®g- Three methods are used for visualizing and measuring ure-eight twin gyre shown in Figs. 4c,d (␦ ϭ 1.25 and the horizontal circulation. The ®rst two involve the 1.4). [The hyperbolic stagnation point is marked by an tracking of neutrally buoyant plastic particles suspended isolated white dot to the lower right of the northern in the tank and illuminated by a horizontal sheet of light (left-hand) recirculation in Fig. 4c.] The twin-gyre re- generated by an argon ion laser. Streak images showing gime persists until ␦ ഡ 1.40 beyond which a single gyre the trajectories of the particles over short time intervals reappears and natural unsteadiness increases, the latter are made from digital video recordings of the ¯ow. Di- a possible result of the interior instability discussed by rect velocity measurements are also made in a limited Meacham and Berloff (1997). The remainder of dis- area using particle image velocimetry (PIV; Raffel et cussion is con®ned to the single (␦ Ͻ 1.1) and twin (1.1 al. 1998). This process involves calculation of the ve- locity ®eld implied by the difference in particle locations 2 Grif®ths and Kiss (1999) ®nd very slow time dependence in a between two video frames taken in close succession. similar experiment, even for values of ␦ Ͻ 1. The camera and analysis algorithm allows for measure- 3 In a similar experiment, Beardsley (1969) notes twin gyre for- ment of velocities with a spatial resolution of 1 cm. As mation at ␦ ϭ 0.75. 1876 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

FIG. 4. Streak images showing nearly steady ¯ow for various ␦. North is to the left and west is downward.

Ͻ ␦ Ͻ 1.4) gyres and a comparison of stirring and that the southern gyre is thinner and less distinct than mixing between these geometries under time-dependent the northern gyre, a feature that has led us to focus perturbations forced by the lid. further investigation on the northern gyre. In order to force the simplest possible time depen- dence, the period T of lid oscillation was chosen to 4. Visualization of stirring, mixing, unstable osc match the period 2␲/⌬⍀ of differential lid rotation (20± manifolds, and turnstile lobes 131 s). This choice insured that any periodic distur- Different Lagrangian aspects of the ¯ow can be vi- bances introduced by imperfections in the lid would sualized by injecting dye into the tank through a needle. simply add to the periodic disturbance forced by the The needle is orientated radially, as shown in Fig. 5, oscillating surface stress. To get a measure of the actual and can be extended or retracted allowing injection into time dependence, we calculated the power spectrum of the southern recirculation interior or the western bound- kinetic energy (Fig. 6) at a location near the eastern ary layer. The approximate boundary of the steady, twin edge of the northern member of the twin gyre for ␦ ϭ recirculation gyre (with ␦ ϭ 1.25) can be visualized by 1.25 and Aosc ϭ 0.15. The velocity at this location was positioning the tip of the needle very close to the west- obtained from PIV measurements over a time span of ern edge of the southern gyre. The resulting dye streak 9 Tosc. The peak near 0.01 Hz corresponds to motions is shown 9 and 21 min after injection is started (Fig. at the forcing frequency and lies an order of magnitude 5). The dye initially moves northward until the hyper- higher in power density than all other peaks save the bolic stagnation point is reached. There it splits into two one between 0.3 and 0.35 Hz. The latter corresponds to streaks, one of which traces the western edge of the the 3.14-s rotation period of the tank and is thought to northern gyre and one of which traces the eastern edge be caused by variations of light intensity at this period of the southern gyre (Fig. 5a). Later the remaining due to imperfections in the rotating ®ber optic coupling.4 boundaries of the gyres have been completely traced, revealing the ®gure-eight separatrix (Fig. 5b). The wisp- 4 The PIV algorithm computes ¯uid velocities by comparing pat- iness of the dye is thought to be due to double diffusion terns set up by (light) beads and (dark) ¯uid at slightly different between the salty ambient water and the food color dye. times. The algorithm works best if the background light intensity is (Salt was added to match the density of the dye.) Note steady and uniform. If the light is not spatially uniform, temporal JUNE 2002 DEESE ET AL. 1877

FIG. 5. Dye intrusion tracing the edge of the steady ®gure-eight gyre. The dimensionless times

t/Tosc ϭ 0 and 5.5 correspond to 9 and 21 min after injection is begun. North is to the left and west is downward. The dye needle is shown as a thin line protruding through the western boundary.

There may also be unsteadiness in the ¯ow at this high to very diffuse distributions. In the single-gyre cases frequency due to irregularities in the drive train, but no (top two frames) the dye remains con®ned to the interior obvious indication of this was detected in the dye vi- of the recirculation, even though the ¯ow is unsteady. sualizations described below. The twin-gyre case (lower frame) produces a more ex- As expected, the time dependence introduces consid- tensive pattern of folds and ®laments that extend east- erably more stirring in the twin-gyre case than the sin- ward away from the gyres. gle-gyre case. One basis for comparison is the dye pat- We believe that the extended strands of dye that terns produced after many forcing periods (Fig. 7, with appear in Fig. 7c result from stretching and folding Aosc ϭ 0.05). The three images show two single-gyre associated with a distinguished hyperbolic trajectory cases (␦ ϭ 0.75 and 1.00) and one twin-gyre case (␦ ϭ (DHT) and its stable and unstable manifolds. Earlier, 1.25). Dye was injected for 23 min in what was judged we sketched these processes by showing the evolution to be the interior of the single recirculation or, in the of lobes of ¯uid trapped between stable and unstable ®nal case, the interior of the southern gyre. The images manifolds in a hypothetical ¯ow (Fig. 2). Very similar were recorded 54 min after injection was started leading objects can be seen in images taken sooner after in- troduction of the dye. The sequence in Fig. 8 shows the evolution of the dye contour over two periods of variations in intensity can induce apparent pattern shifts, causing the the lid oscillation for the twin gyre with ␦ ϭ 1.25, algorithm to detect false velocities. Aosc ϭ 0.05, and Tosc ϭ 131 s. As before, dye is in- 1878 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

FIG. 6. Power spectrum of ¯uid speed calculated at a location in the eastern portion of the northern recirculation. Dashed lines indicate 95% con®dence limits.

jected very close to the western boundary of the we see that two additional southward ®laments have southern gyre, just upstream of the suspected hyper- formed. Precise measurements indicate that Tosc is in- bolic trajectory. The hope was that some of the dye deed the time interval between passage of the edges would ®nd its way into the region around the hyper- of consecutive ®laments across a ®xed location. All bolic trajectory and, from there, split and follow the of these features are characteristic of the unstable unstable manifolds. About three cycles after injection manifold for the northern gyre of the hypothetical is begun (Fig. 8a, t/Tosc ϭ 0), the initial dye line has ¯ow (Fig. 2c). moved along the western edge of the southern gyre The above picture suggests an exchange process be- and split at the hyperbolic trajectory, forming a north- tween the northern gyre and surrounding ¯uid. A quan- ward (to the left) moving strand that encircles the titative description of this transport could be made by northern gyre and a southern strand that moves along locating the stable manifolds of the DHT, allowing the east side of the southern gyre. The northern strand turnstile lobes to be identi®ed. Although stable man- does not close to form the northern boundary of the ifolds can be calculated in numerical simulations by ®gure-eight, as it does in the steady case. If fact, the running time backward, no simple method for ®nding tip of the strand can be seen in Fig. 8b far to the right a stable manifold in a laboratory experiment exists. In of the DHT. Beginning at this tip and following the principle, one could store PIV velocity information on strand backward (to the left), one encounters several a horizontal grid over a full period of oscillation and, sharp turns or meanders. The ®rst sharp turn is part with this complete velocity record, do a backward time of a narrow wedge of ¯uid that is beginning to wrap simulation. Instead, we have attempted to identify turn- around the inside edge of the northern gyre. As time stile lobes by approximating the position of the stable progresses, additional meanders form in the strand at manifold extending immediately northeastward from the eastern edge of the northern gyre (Fig. 8c). These the hyperbolic trajectory. That is, we approximate meanders steepen and fold into wedges or ®laments something analogous to the solid line in Fig. 2c ex- (Fig. 8g) that are alternately pulled into or drift south- tending a short distance to the upper right of the DHT. ward of the northern gyre. The ®laments pulled into The approximation consists of a short line segment the northern gyre spiral anticyclonically about the in- extending northeastward from the suspected position side edge. The southward ®laments continue to drift of the hyperbolic trajectory at a given time and in a southward along the eastern side of the southern gyre. direction that bisects the ®laments so as to give equal

After nearly two cycles (t/Tosc ϭ 1.8) of lid oscillation, area on either side of the line. As described in section JUNE 2002 DEESE ET AL. 1879

FIG. 7. Dye patches 54 min after injection for unsteady ¯ows (Aosc ϭ 0.05) with single-gyre geometry (␦ ϭ 0.75, 1.00) and twin-gyre geometry ␦ ϭ 1.25. North is to the left and west is downward. 1880 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

FIG. 8. Evolution of dye lines over two forcing cycles for ␦ ϭ 1.25 and Aosc ϭ 0.05. North is to the left and west is downward. JUNE 2002 DEESE ET AL. 1881

FIG. 9. Dye lines similar to those shown in Fig. 8 with an approximation (the straight line segment) to the stable manifold for the northern gyre. The darkened patches are approximations

of turnstile lobes; their areas are labeled above each frame: (a)±(c) ␦ ϭ 1.25 and Aosc ϭ 0.05,

(d)±(f) ␦ ϭ 1.25 and Aosc ϭ 0.15. North is to the left.

2, the requirement of equal lobe areas, a consequence there are uncertainties involved with this exercise. The of mass conservation, holds as long as the horizontal position of the hyperbolic trajectory must be guessed velocity is nondivergent, the time-dependence is pe- and the use of a straight, rather than curved line, means riodic, and only a single lobe enters the recirculation that the three lobe areas can almost never be exactly during each period.5 In the experiment the ®rst two equal. Nevertheless, the construction allows one to vi- conditions hold to within a good approximation where- sualize approximations to the real lobes. The darkened as the third holds exactly. lobe in frame (a) contains ¯uid entering the recirculation We have applied the above method in order to ap- while those of (b) and (c) contain ¯uid that is leaving. proximate a short section of the stable manifold for the Likewise, the frame (d) darkened lobe contains entering Fig. 8 ¯ow. The result is a hand-drawn line segment ¯uid while the frames (e) and (f) darkened lobes contain extending from the position of the DHT northeastward departing ¯uid. (Figs. 9a±c). It intersects the wavy unstable manifold The ¯ux or ``chaotic transport'' F of ¯uid into (or such that the ®rst three identi®able lobes have approx- out of) the northern gyre can be calculated by dividing imately equal areas. Each of the three lobes is darkened the area of any single lobe by the period Tosc (section individually in Figs. 9a±c and their areas are given. 2). We have carried out this calculation using the av- When Aosc is increased from 0.05 to 0.15 the corre- erage of the lobe areas shown in Figs. 9a±c (or 9d±f) sponding lobes become larger (Figs. 9d±f). Naturally, as an estimate of the actual lobe area. From the resulting F, a ¯ushing time for the northern gyre can be estimated 5 In some cases, a requirement of equal lobe volume is appropriate. by de®ning the gyre boundary as the union of the line For example, if the bottom slope in the tank was large enough to segment in Fig. 9 with the portion of the unstable man- cause signi®cant depth differences on the scale of the recirculation, ifold extending from the northern end of the line seg- and if the ¯ow obeyed inviscid shallow-water dynamics [implying ment counterclockwise back to the other end of the line. z ϭ 0], then the manifolds wouldץ/␷ץz ϭץ/uץ y ± 0 butץ/␷ץ x ϩץ/uץ be independent of depth and the lobes would have unequal horizontal (The resulting gyre area Ag is approximately the same area but equal volume. The presence of friction boundary layers at as in the steady case.) Measuring the enclosed area gives the top or bottom might invalidate this equality. ¯ushing times Ag/F ഡ 9Tosc for the case of weaker time 1882 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

dependence (Aosc ϭ 0.05) and about 3Tosc for Aosc ϭ of passive tracer concentration, here the dye gray scale. 0.15. Interpretation of ¯ushing is complicated by the This area may be noncontiguous and the total A includes fact that the lobes do not appear to ®ll the entire area all ``islands'' in which the tracer concentration c is great- inside the recirculation, as suggested by the clear area er than or equal to C. Because A(C, t) is a single-valued in the center of the northern gyre in Fig. 7c. Thus the function, the inverse function C(A, t) can also be de- actual area being ¯ushed is probably smaller than Ag. ®ned. If the advection±diffusion equation Anthropogenic ocean tracers tend to be introduced at cץ (the source region of a deep western boundary current ϩ J[␺, c] ϭ ␬ٌ2c ϩ S (1 tץ and are thought to be transported from the boundary current into the recirculations. The fate of a tracer in- is integrated over A(C, t) and the condition of incom- troduced in the western boundary layer of the laboratory pressibility is used, the resulting constraint can be writ- ¯ow is shown in Fig. 10a. Dye is injected just outside ten, after some manipulation, as of the southern gyre for a time-dependent case with ␦ ץץ (C(A, tץ ϭ 1.25, Aosc ϭ 0.05. The dye is carried northward and 2 eventually around the northern and eastern edges of the ϭ ␬LCe (A, t) . (2) Aץ[]Aץ tץ northern gyre. Some of this dye is transported into and later expelled from the northern gyre by a process re- The source term S in (1) has been neglected as we con- sembling the turnstile mechanism. Evidence for the lat- sider tracer distributions only after injection has been 2 ter exists in the form of alternating dark and light ®l- terminated. The diffusion coef®cient ␬eff ϭ ␬Le in (2) is aments that pass a ®xed location at exactly the forcing the product of the molecular or subgrid-scale diffusivity period. However, the dye has not been introduced so as ␬ and the square of the effective length, de®ned by to map out the manifolds and the ®laments shown are ١c|2͘|͗ therefore not exactly turnstile lobes. Some of the ma- L2 ϭ . (3) A)2ץ/Cץ) terial expelled from the recirculation penetrates far east- e ward into the Sverdrup interior. The operator ͗(´)͘ denotes average over the area be- Tracer injected just inside the western edge of the tween adjacent tracer contours and is de®ned by southern gyre evolves in a qualitatively similar way (Fig. 10b). Some of the dye makes its way into the dl northern gyre and escapes in ®laments similar to the (´) |١c| above. Only when dye is injected into the very center Ͷ ͗(´)͘ϭ or, equivalently (4a) of the southern gyre does it remain largely con®ned dl |١c| there (Fig. 10c). This con®nement is an indication that the center of the gyre is protected by a barrier penetrable Ͷ ץ only by molecular diffusion. Evidence for a barrier in the interior of the northern gyre can be seen in the form ͗(´)͘ϭ (´) dA. (4b) Aץ of dye-free centers in Figs. 10a and 10b (also see Figs. ͵͵ 7c and 8). In essence, (1) has been reduced to a diffusion equa- tion in A-space in which the effect of advection is con- tained in the diffusion coef®cient ␬L2. The effective 5. Effective diffusivity calculations e length Le is a measure the convolution of the tracer a. Background contours, and will grow as the tracer is stirred. Here Le is not the actual length L of the contour; however, it

Although the visual evidence is that the ®gure-eight can be shown (Haynes and Shuckburgh 2000a) that Le geometry of the twin gyre leads to considerably more Ͼ L. When ␬eff is plotted as a function of A (or C), stirring and mixing under time dependence than the sin- minima generally correspond to barriers to mixing gle-gyre geometry, this remains to be quanti®ed. To do whereas maxima correspond to regions of active trans- so, we employ the theory of effective diffusivity (Nak- port and mixing. These relationships are most mean- amura 1996; Winters and D'Asaro 1996), which pro- ingful if ␬eff is independent of the initial tracer ®eld, a vides for bulk quanti®cation of the catalytic effect of situation that can often be realized by allowing the tracer stirring in producing molecular (or subgrid scale) prop- to spread and mix to the point where it is uniformly erty exchanges. Stirring is indicated by the geometric distributed along, but not across, the time-average complexity of the ¯ow resulting from the contortion of streamlines of the ¯ow. Haynes and Shuckburgh tracer contours due to advection. The product of the (2000a,b) note that an adjustment period of about a analysis is an ``effective'' diffusion coef®cient ␬eff that month is required in simulations of the stratosphere. depends on both the contortion in the tracer ®eld and Generally speaking, the magnitude of ␬eff is interesting the molecular diffusivity. only in comparison between different experiments. The formal theory involves de®nition of a new var- Since ␬ is constant in our laboratory tank, further dis- iable A(C, t), the area enclosed by a given contour C cussion is limited to comparisons of Le distributions. JUNE 2002 DEESE ET AL. 1883

FIG. 10. Dye patterns found for the case ␦ ϭ 1.25 and Aosc ϭ 0.05 with (a) injection in the western boundary current just to the west of the southern gyre; (b) injection just inside the edge of the southern gyre; (c) injection well inside the southern gyre. North is to the left. 1884 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

Our use of effective diffusivity is a bit different from c. Results what has just been described. We wish to compare the magnitude of mixing in a bulk sense between time- Chaotic advection sets in above the transition from dependent single- and double-recirculation gyres. The the unsteady single-gyre to the unsteady twin-gyre. To quantify the preferential stirring claimed to be due to presence of hyperbolicity in the latter but not the former chaotic advection, it necessary to show ®rst that this suggest that L should be noticeably higher on average e increase is not due to the geometric transition alone in the latter. Calculations of L are made at times shorter6 e (without time dependence). We do so by calculating than the adjustment period suggested above, and the 2 Le(C) for a range of steady cases with increasing ␦. The results are therefore sensitive to how the dye is intro- single- and twin-gyre cases ␦ ϭ 1.00 and ␦ ϭ 1.25 have duced. Comparisons between experiments were there- 2 very similar, if noisy,Le (C) (Fig. 11a). The results are fore made using identical injection times and rates, and based on images taken 31 min after injection was halted. by maintaining the injection point at a ®xed location By contrast, the introduction of time dependence (with (or, in some cases, at what was judged to be ®xed rel- Aosc ϭ 0.05) leads to signi®cant differences (Fig. 11b). ative to the edge of the recirculation). The experimen- An order of magnitude increase in the mean and peak 2 2 tally determinedLe (C) is typically noisy compared to values ofLe occurs in the twin gyre (␦ ϭ 1.25) case. that obtained in numerical simulations (e.g., Nakamura It can also be shown that increasing inertia in the 1996; Haynes and Shuckburgh (2000a,b) where the ini- unsteady singe gyre does not signi®cantly increase the 2 tial tracer distribution is smooth and where a longer stirring. When the mean values ofLe are calculated over adjustment period is allowed for. The noise level of our a range of ␦, and with Aosc ϭ 0.05 in each case, the distributions is large enough to make identi®cation of resulting values (Fig. 12) are remarkably similar for all barriers dif®cult, but weak enough to render compari- single-gyre cases (␦ Ͻ 1.10). Frames (a) and (b) show sons of different experiments meaningful. results based on data measured 32 and 54 min after injection is begun (and 9 min and 21 min after injection is terminated). Open circles indicate an injection point b. Procedure just inside the western or eastern edge of the southern gyre, while asterisks indicate injection well into the in- Effective diffusivity can be calculated from a snap- terior of the southern gyre. The setting ␦ ϭ 1.12, which shot of the tracer (dye) ®eld and the calculation itself is considered borderline between the two geometries, includes only integrals and derivatives with respect to has similar values to those with lower ␦. A signi®cant 2 the area A within contours if the de®nition (4b) is used. jump in meanLe generally occurs upon transition to the The function A(C) is constructed by counting for each twin gyre geometry (␦ ϭ 1.25), particularly for dye C the number of gray-scale pixel values greater than C injected just inside the edge of the southern gyre (open and multiplying by the area-to-pixel ratio. In addition, circles). Increased sensitivity to the injection point is a|ٌc | 2 value is estimated for each pixel using a second- also apparent at ␦ ϭ 1.25, but all cases with injection order centered difference approximation. These values near the gyre edge are substantially larger than those are summed to form the area integral in the expression for lower ␦. ,A ## | ٌc | 2 dA. No contouring of Cץ/ץfor ͗ | ٌc | 2͘ϭ nor calculation of line integrals is required. 6. Discussion Dye is typically injected into the tank at a constant 2 Chaotic advection is a possible mechanism for ¯uid rate at a ®xed location for 23 min. AllLe calculations are from images taken after the dye pump is stopped. exchange and mixing among a western boundary current In all cases the adjustment time is longer than the typical and adjacent recirculation gyres. Geometries favorable winding time (2±4 min in the northern gyre and 10 min for this process contain a distinguished hyperbolic tra- in the southern gyre) of particles within recirculations. jectory or at least a region of strong and persistent Although some vertical mixing and diffusion of the dye stretching and contraction, such as the self- intersection occurs, the visual evidence is that the dye became ver- of our ®gure-eight gyre. In the presence of time-depen- tically homogenized within a 5±10 cm-thick layer about dence, trajectories brought into close contact with a the level of the needle during the injection and adjust- DHT tend to become chaotic. Lagrangian chaos in such ment time. a setting is limited to regions occupied by the folded and tangled stable and unstable manifolds of the DHT. Associated transport processes can be quanti®ed by identifying turnstile lobes whose boundaries are de®ned 6 Practical considerations place a limit on the adjustment time for by sections of the intersecting manifolds. We have at- the dye. As the dye spreads out, it tends to wander into reaches of tempted to visualize some of these objects through care- the tank not easily covered by our imaging system. In order to measure A(C), it is necessary to have an image in which all C contours are ful injection of dye in a laboratory model. The spreading closed, which is ensured when the dye has not spread to the edge of of dye in the twin-gyre geometry has been compared the image. to a single-gyre case that has no region of strong hy- JUNE 2002 DEESE ET AL. 1885

2 FIG. 11. (a)Le (C) for steady single (␦ ϭ 1.00) and twin (␦ ϭ 1.25) recirculation gyres. (b) 2 Le(C) for the same cases but with unsteady forcing (Aosc ϭ 0.05). perbolicity and is therefore unfavorable for chaotic ad- can be visualized (Fig. 9) and the associated chaotic vection. Quantitative comparisons of mixing between transport and ¯ushing time estimated. The dye patterns the single- and twin-gyre cases have been made based also indicate barriers to transport in the middle of the on the effective diffusivity calculated from snapshots of gyres, as evidenced by the clear region in the northern the dye ®elds. gyre (Fig. 7c or 10a,b) and the trapped dye in the center Both the visual and quantitative evidence suggests of the southern gyre (Fig. 10c). Long-term snapshots of the presence of preferential exchange, stirring, and mix- the dye (e.g., Fig. 10) suggest how tracers are trans- ing in the twin-gyre case. The process by which western ported into the interior basin circulation. Measurements boundary layer ¯uid is advected into, and later expelled of the effective length Le (that appears in the de®nition 2 from, the northern member of the twin gyre is suggested of effective diffusivity ␬eff ϭ ␬Le ) con®rm preferential by visualization of the unstable manifold (or something stirring in the twin-gyre case as opposed to the single- 2 very close to it) of the northern gyre (Fig. 8). Also, by gyre case. Unfortunately, the distributions ofLe as a approximating a segment of the stable manifold using function of tracer concentration C are suf®ciently noisy the principle of equal lobe areas, some of the lobes that to preclude identi®cation of barriers, as has been done deliver and export ¯uid to and from the northern gyre in previous numerical studies. 1886 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

2 2 FIG. 12. Mean values ofLe (cm ) in unsteady cases presented as a function of ␦ ϭ ␦I/␦M and

calculated at (a) t ϭ 32 min and (b) 54 min. In all cases Aosc ϭ 0.05. Open circles represent cases with dye injected into the western or eastern part of the southern recirculation and asterisks represent cases with dye injected into the very center of the southern recirculation. Bars represent qualitative estimates of uncertainty based on cases in which runs were repeated several times.

Examples of chaotic advection are often characterized a hyperbolic region that exponential separation occurs. by a separation between Lagrangian and Eulerian time- This picture provides a connection with conventional scales, with TL K TE. In fact, this scale separation is chaotic behavior, a connection that would become more sometimes taken as one of the de®ning characteristics tenuous, and perhaps lost, if the hyperbolic regions were of chaotic advection (Shepherd et al. 2000). An inter- to rapidly disappear and reform. The requirement TL K pretation of this requirement is that the local hyperbol- TE is therefore made under the expectation that the La- icity underpinning Lagrangian chaos must persist over grangian motion will live in a slowly varying Eulerian many ``winding'' times of the particle motion. A pair velocity ®eld and that, in particular, the hyperbolic re- of closely spaced parcels will therefore travel through gions will be long lived. Shepherd et al. (2000) argue the ¯ow ®eld, occasionally wandering close enough to that the stratosphere (a region where chaotic advection JUNE 2002 DEESE ET AL. 1887

is thought to be relevant) is characterized by TL K TE, just this using a numerical model of the Gulf of Mexico whereas the mesosphere (where chaotic advection is a circulation.] However, proof that chaotic advection is poor model for stirring) is characterized by TL ഡ TE. relevant requires more than just formal identi®cation of It is also possible, however, for persistent hyperbolic these objects. The organizing structures and their hy- structures to exist in ¯ows where TL ഡ TE. Examples perbolic trajectories must be persistent. Moreover, the can be arranged by tuning parameters in simple models manifolds must be distinct and remain smooth on the (e.g. Bower 1991; Samelson 1992; Pratt et al. 1995). scale of the organizing structures, indicating that ad-

Though TL ഡ TE in our experiment, the ¯ow is appar- vection by these structures, and not by smaller eddies, ently characterized by persistent strong hyperbolicity in is responsible for the stretching and folding of ¯uid the twin-gyre con®guration. Thus, many of the Lagrang- elements and the generation of tracer ®ne structure. In ian objects that typify chaotic advection (manifolds, contrast, the stretching and folding in a conventional lobes, barriers, etc.) can still be found. 2D turbulent enstrophy or energy cascade is strongly Although visual evidence (e.g., Figs. 10a,b) suggests in¯uenced by a range of scales. Dye injected at a that exchange between the gyres and their surroundings ®xed location evolves into a plume, rather than the occurs to the east of the gyres, such a judgment is en- smooth ®laments seen in Figs. 8 and 9. Hyperbolic tra- tirely subjective. A boundary for the recirculation can jectories are intermittent and their stable and unstable be de®ned as the union of segments of the stable and manifolds are indistinct and limited in length. In sum- unstable manifolds (Fig. 2c or 2d). The intersection mary, the main challenge in establishing chaotic ad- point at which the segments are joined can be chosen vection as an oceanographically relevant process is not in many ways and one's interpretation of where ¯uid in identifying organizing structures, hyperbolic trajec- enters and leaves the recirculation depends on this tories, and the like, but rather in eliminating turbulence choice. For example, joining the stable and unstable as the mechanism controlling transport and stirring. manifolds at the intersection between lobes AЈ and B in Shepherd et al. (2000) discuss ways in which atmo- Fig. 2c would formally lead to exchange across the west- spheric data has been used to distinguish between cha- ern side of the gyre. In the laboratory gyre, the lobes otic advection and conventional turbulence. Wavenum- at the western side are undoubtedly narrow and elon- ber spectra provide a major basis for comparison. Since gated, a feature associated with the relatively high ¯uid small-scale eddies are relatively unimportant in chaotic speeds in this region. The lobes here are de®ned pri- advection, the corresponding kinetic energy or passive marily by the highly convoluted stable manifold, which tracer spectra tend to decay more steeply than those for is hidden in our images. On the eastern side of the gyre, classical turbulence. In the oceanographic context, such where the stable manifold has relatively little curvature, a comparison would require wavenumber spectra of the lobes are de®ned primarily by the highly convoluted these quantities along isopycnal surfaces. Such products and easily visualized unstable manifold. The resulting are virtually unknown at present. Stammer (1997) has picture is one in which the lobes are more apparent at calculated global wavenumber spectra of kinetic energy the eastern side of the gyre, giving the impression that based on satellite altimeter measurements. Over the exchange occurs there. A sound way of thinking about range of resolvable scales of a few thousand to about the exchange is that its location depends on how the 50 km, only the band 50±200 km indicates a decay steep boundary is de®ned but that its magnitude does not. enough to be a candidate for chaotic advection. However Having said this, it is also apparent from the experiments these data are based on measurements of sea surface that the zonal spreading of a tracer away from the west- height and do not directly relate to quasi-2D motion ern boundary is most rapid in the region to the east of along particular isopycnal surfaces. Also, scales below the twin gyres. the deformation radius, one of the most likely ranges The present work has suggested chaotic advection as of relevance for chaotic advection, are not resolved. a possible mechanism for transport and mixing within More promising discriminants might follow from the and between boundary currents and recirculations. analysis of ¯oat and drifter data. The stirring associated However, the relevance of this mechanism in the ocean with chaotic advection is much less homogeneous than has not been established. In the search through spatial that produced by conventional turbulence and this scales over which chaotic advection might be important, should show up in the statistics of dispersion. In prin- the focus should be on motions that are quasi-two-di- ciple, a pair of closely spaced ¯oats launched in the mensional, either horizontally or along isopycnals. Giv- (nonchaotic) center of either of the twin gyres will sep- en an isopycnal velocity ®eld containing mesoscale fea- arate from each other at a rate linear in time. A similar tures, say, it would probably not be hard to ®nd coherent pair launched in the (chaotic) region of tangled mani- eddies or recirculations that might serve as the orga- folds should experience separation at a rate exponential nizing structures for chaotic advection. Further, it should in time. As shown by Shepherd et al. (2000) the presence be possible to locate hyperbolic trajectories within such of distinct behavior in different regions of the ¯ow ®eld ¯ow ®elds and to plot their stable and unstable mani- causes PDFs of measures of relative dispersion to be folds, at least over ®nite spans in time. [Kuznetsov et highly skewed. In conventional turbulence, which lacks al. (2000, submitted to J. Phys. Oceanogr.) have done barriers and islands of nonchaotic motion, ¯oats are free 1888 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32 to sample the entire ¯ow ®eld and the resulting PDFs d. Dye injection are much less skewed. To our knowledge, this idea has We used McCormick red food coloring 1.02391 not been exploited in the analysis of ocean ¯oat and ␳ ϭ (kg mϪ3) diluted to match the 1.022 (kg mϪ3) tank drifter data. ␳ ഡ water at 20ЊC. This mixture sat in a reservoir (on the rotating table) and was pumped through tygon tubing Acknowledgments. Deese was supported during the and into the feed needle by a variable ¯ow, peristaltic course of this work by a National Defense Science and pump. This Variable Flow Mini Pump fed the dye ¯ow Engineering Fellowship from the U.S. Department of through a section of 0.14-cm diameter tubing at a rotor Defense. Pratt was supported by a grant from the Of®ce rpm ϳ0.5. This results in a calculated ¯ow rate of ap- of Naval Research (N00014-99-1-0258) and Helfrich by proximately 6 ϫ 10Ϫ3 cm sϪ1, which would result in a a grant from the National Science Foundation (OCE- total input of 8 cm3 over the course of each run. As 9616949). The authors wish to thank R. Beardsley, C. Figs. 5, 8, and 9 show, dye leaves the needle and enters Cenedese, C. Mauritzen, M. McCartney, Robert Pickart, the ¯uid at nearly a right angle to the needle (approx- A. Rogerson, J. Salzig, F. Straneo, J. Wells, and two imately parallel to the western boundary current ¯ow) anonymous reviewers for helpful suggestions and as- implying minimal disturbance of the ambient velocity sistance that was given during the course of this work. ®eld by the dye stream. APPENDIX e. PIV analysis and lobe dynamics Technical Information In principle, PIV velocities could be obtained over the entire tank at regular time slices. If the ¯ow were a. Video and laser exactly periodic in time, data acquired over one period The Pulnix video camera was attached to a frame would provide gridded velocities for all space and time, rotating with the table so that images are stationary with raising the possibility of performing a lobe analysis. respect to the tank. The neutrally buoyant particles were Miller et al. (1977) carried out a similar analysis using illuminated using a 6 W argon ion laser mounted away nearly periodic data obtained from a numerical model from the rotating table. The laser beam was transmitted of a meandering jet. The lack of exact time periodicity through a ®ber-optic cable and rotating coupling onto leads to small discontinuities at the end of each nominal the table. A specialized lens brought the beam into a time period. Willingness to live with the associated er- horizontal sheet of light, which was shone through the rors allows turnstile lobes to be tracked over several tank about 5 cm below the lid. To visualize dye, two periods. The lobes tend to be stretched into ®ne ®la- 20 W ¯uorescent lightbulbs were positioned facing up- ments whose width rapidly decreases to below that af- ward below the tank. Blackout cloths were suspended forded by the numerical resolution. from the ceiling and pinned around the tank to cut down The observed lobes have dimensions of 1±2 cm (Fig. as much as possible on background light. Images were 9). The darkened lobe in frame a evolves over one forc- recorded on a VCR and digitized using Mutech Mv- ing period (Tosc) into something close to the darkened 1000 digital imaging software. lobe in frame c. After the next period, this lobe is then stretched at a rate that causes its width to decrease to

a fraction of a centimeter over the next Tosc (Fig. 8). b. Neutrally buoyant particles Lobes that are entering the northern gyre (e.g., Fig. 9a) In order to match the density of the particles to the presumably evolve from very thin lobes carried north density of the liquid in the tank, we used Pliolite plastic by the western boundary layer. Some of this ¯uid may particles (␳ ϭ 1.024 kg mϪ3) and saltwater (␳ ഡ 1.022 have been expelled from the southern gyre. It would be kg mϪ3) ®ltered at 50 ␮m. These particles were hand interesting to follow all of these lobes backward and ground with a mortar and pestle and then sorted into forward in time to learn more about the origin and fate size classes using a series of sieves. Particles of diameter of ¯uid that is carried into and out of the northern gyre. d Ն 250 ␮m were chosen to make streak diagrams and Unfortunately, the camera resolution and PIV algorithm particles of diameter 150 ␮m Յ d Յ 250 ␮m were allows velocities to be measured with a spatial resolu- chosen for PIV analysis. tion of at best 1 cm, lower than required to resolve the highly ®lamented lobes. c. Spinup REFERENCES Twelve minutes of spinup time was allowed for all runs. This time was much longer than the viscous spinup Beardsley, R. C., 1969: A laboratory model of the wind-driven ocean Ϫ1/2 Ϫ1 circulation. J. Fluid Mech., 38, 255±271. time E k f ഡ 60 s and the ഡ30 s time required for Bower, A. S., 1991: A simple kinematic mechanism for mixing ¯uid the fastest Rossby wave to cross the tank. Unsteadiness parcels across a meandering jet. J. Phys. Oceanogr., 21, 173± in the ¯ow was not visually observed after 7 or 8 min. 180. JUNE 2002 DEESE ET AL. 1889

ÐÐ, and H. D. Hunt, 2000a: Lagrangian observations of the Deep boundary current of the northern North Atlantic. Progress in Western Boundary Current in the North Atlantic Ocean. Part I: Oceanography, Vol. 29, Pergamon Press, 283±383. Large-scale pathways and spreading rates. J. Phys. Oceanogr., Meacham, S. P., and P. S. Berloff, 1997: Instabilities of a steady, 30, 764±783. barotropic, wind-driven circulation. J. Mar. Res., 55, 885±913. ÐÐ, and ÐÐ, 2000b: Lagrangian observations of the Deep Western Miller, P. D., C. K. R. T. Jones, A. M. Rogerson, and L. J. Pratt, 1997: Boundary Current in the North Atlantic Ocean. Part II: The Gulf Quantifying transport in numerically-generated velocity ®elds. Stream±Deep Western Boundary Current crossover. J. Phys. Physica D, 110, 105±122. Oceanogr., 30, 784±804. ÐÐ, L. J. Pratt, K. R. Helfrich, and C. K. R. T. Jones, 2002: Chaotic Deese, H. E., 2001: Chaotic advection and mixing in a Western transport of mass and potential vorticity for an island recircu- Boundary Current±recirculation system: Laboratory experi- lation. J. Phys. Oceanogr., 32, 80±102. ments. M.S. thesis, Woods Hole Oceanographic Institution±MIT Molinari, R. L., R. A. Fine, W. D. Wilson, R. G. Curry, J. Abell, and Joint Program, 118 pp. M. S. McCartney, 1998: The arrival of recently formed Labrador Fine, R. A., 1995: Tracers, time scales and the thermohaline circu- Sea Water in the Deep Western Boundary Current at 26.5ЊN. lation: The lower limb in the North Atlantic Ocean. Rev. Geo- Geophys. Res. Lett., 25, 2249±2252. phys., 33 (Suppl.), 1353±1365. Nakamura, N., 1996: Two-dimensional mixing, edge formation, and Friedrichs, A. M., and M. M. Hall, 1993: Deep circulation in the permeability diagnosed in an area coordinate. J. Atmos. Sci., 53, tropical North Atlantic. J. Mar. Res., 51, 697±736. 1524±1537. Grif®ths, R. W., and A. E. Kiss, 1999: Flow regimes in a wide ``sliced- ÐÐ, and J. Ma, 1997: Modi®ed Lagrangian-mean diagnostics of the cylinder'' model of homogeneous beta-plane circulation. J. Fluid stratospheric polar vortices. Part 2: Nitrous oxide and seasonal Mech., 399, 205±236. barrier migration in the cryogenic limb array etalon spectrometer Haynes, P., 1999: Transport, stirring, and mixing in the atmosphere. and SKYHI general circulation model. J. Geophys. Res., 102 (D22), 25 721±25 735. Chaos and Turbulence, H. ChateÂ, E. Villermaux, and J.-M. Pedlosky, J., and H. P. Greenspan, 1967: A simple laboratory model Chomaz, Eds., Kluwer, 406 pp. for the oceanic circulation. J. Fluid Mech., 27, 291±304. ÐÐ, and E. Shuckburgh, 2000a: Effective diffusivity as a diagnostic Pickart, R. S., and N. G. Hogg, 1989: A tracer study of the deep Gulf of atmospheric transport. Part I: Stratosphere. J. Geophys. Res., Stream cyclonic recirculation. Deep-Sea Res., 36, 935±956. 105 (D18), 22 777±22 794. ÐÐ, ÐÐ, and W. M. Smethie Jr., 1989: Determining the strength ÐÐ, and ÐÐ, 2000b: Effective diffusivity as a diagnostic of at- of the Deep Western Boundary Current using the Chloro¯uo- mospheric transport. Part II: Troposphere and lower stratosphere. romethane ratio. J. Phys. Oceanogr., 19, 940±951. J. Geophys. Res., 105 (D18), 22 795±22 810. Pratt, L. J., M. S. Lozier, and N. Beliakova, 1995: Parcel trajectories Hogg, N. G., 1992: On the transport of the Gulf Stream between in quasigeostrophic jets: Neutral modes. J. Phys. Oceanogr., 25, Cape Hatteras and the Grand Banks. Deep-Sea Res., 39, 1231± 1451±1466. 1246. Raffel, M., C. Willert, and J. Kompenhans, 1998: Particle Image Ide, K., F. Lekien, and S. Wiggins, 2001: Distinguished hyperbolic Velocimetry: A Practical Guide. Springer-Verlag, 253 pp. trajectories in time dependent ¯uid ¯ows: Analytical and com- Rhein, M., 1994: The Deep Western Boundary Current: Tracers and putational approach for velocity ®elds de®ned as data sets. Non- velocities. Deep-Sea Res., 41, 263±282. linear Proc. Geophys., in press. Rogerson, A. M., P. D. Miller, L. J. Pratt, and C. K. R. T. Jones, 1999: Johns, E., R. A. Fine, and R. L. Molinari, 1997: Deep ¯ow along the Lagrangian motion and ¯uid exchange in a barotropic mean- Western Boundary south of the Blake Bahama Outer Ridge. J. dering jet. J. Phys. Oceanogr., 29, 2635±2655. Phys. Oceanogr., 27, 2187±2208. Rom-Kedar, V., and S. Wiggins, 1990: Transport in two-dimensional Kuznetsov, L., M. Toner, A. D. Kirwan, C. K. R. T. Jones, L. Kantha, maps. J. Arch. Ration. Mech. Anal., 109, 239±298. and J. Choi, 2001: The Loop Current and adjacent rings delin- Samelson, R. M., 1992: Fluid exchange across a meandering jet. J. eated by Lagrangian analysis. J. Phys. Oceanogr., submitted. Phys. Oceanogr., 22, 431±440. Lavender, K. L., R. E. Davis, and W. B. Owens, 2000: Mid-depth Schmitz, W. J., and M. S. McCartney, 1993: On the North Atlantic recirculation observed in the interior Labrador and Irminger Seas Circulation. Rev. Geophys., 31, 29±49. by direct velocity measurements. Nature, 407, 66±69. Shepherd, T. G., J. N. Koshyk, and K. Ngan, 2000: On the nature of Leaman, K. D., and P. S. Vertes, 1996: Topographic in¯uences on large-scale mixing in the stratosphere and mesosphere. J. Geo- recirculation in the Deep Western Boundary Current: Results phys. Res., 105, 12 433±12 446. Stammer, D., 1997: Global characteristics of ocean variability esti- from RAFOS ¯oat trajectories between the Blake±Bahama Outer mated from regional TOPEX/Poseidon altimeter measurements. Ridge and the San Salvador ``gate.'' J. Phys. Oceanogr., 26, J. Phys. Oceanogr., 27, 1743±1769. 941±960. Thompson, L., 1995: The effect of continental rises on the wind- Lee, T. N., W. E. Johns, R. J. Zantopp, and E. R. Fillenbaum, 1996: driven ocean circulation. J. Phys. Oceanogr., 25, 1296±1316. Moored observations of Western Boundary Current variability Watts, D. R., 1991: Equatorward currents in the temperature range and at 26.5ЊN in the subtropical North 1.8Њ±6ЊC on the continental slope in the mid-Atlantic Bight. Atlantic. J. Phys. Oceanogr., 26, 962±982. Proc. Deep Convection and Deep Water Formation Workshop, Malhotra, N., and S. Wiggins, 1998: Geometric structures, lobe dy- Vol. 51, Monterey, CA, Naval Postgraduate School, 183±196. namics, and Lagrangian transport in ¯ows with aperiodic time- Wiggins, S., 1992: Chaotic Transport in Dynamical Systems. Vol. 2, dependence, with applications to Rossby wave ¯ow. J. Nonlinear Interdisciplinary Applied Mathematics, Springer-Verlag, 301 pp. Sci., 8, 401±456. Winters, K., and E. D'Asaro, 1996: Diascalar ¯ux and the rate of McCartney, M. S., 1992: Recirculating components to the deep ¯uid mixing. J. Fluid Mech., 317, 179±193.