Numerical Study of the Generation and Propagation of Trigger Meanders of the Kuroshio South of Japan

Similar documents
The Effect of Koshu Seamount on the Formation of the Kuroshio Large Meander South of Japan

Dynamic Structure of the Kuroshio South of Kyushu in Relation to the Kuroshio Path Variations

Processes Coupling the Upper and Deep Ocean on the Continental Slope

Coastal Disturbance in Sea Level Propagating along the South Coast of Japan and Its Impact on the Kuroshio

Warm Water Intrusion from the Kuroshio into the Coastal Areas South of Japan

On the eddy-kuroshio interaction: Meander formation process

Characteristics of Variations of Water Properties and Density Structure around the Kuroshio in the East China Sea

Short-Range Prediction Experiments with Operational Data Assimilation System for the Kuroshio South of Japan

Scattering of Semidiurnal Internal Kelvin Wave at Step Bottom Topography

Island Wakes in Shallow Water

Annual Variation of the Kuroshio Transport in a Two-Layer Numerical Model with a Ridge

Baroclinic Rossby waves in the ocean: normal modes, phase speeds and instability

Current Variation in the Sea near the Mouth of Suruga Bay*

Mean Stream-Coordinate Structure of the Kuroshio Extension First Meander Trough

General Comment on Lab Reports: v. good + corresponds to a lab report that: has structure (Intro., Method, Results, Discussion, an Abstract would be

Western Boundary Currents. Global Distribution of Western Boundary Currents and their importance

Wind Gyres. curl[τ s τ b ]. (1) We choose the simple, linear bottom stress law derived by linear Ekman theory with constant κ v, viz.

Buoyancy-forced circulations in shallow marginal seas

Vortices in the ocean. Lecture 4 : Baroclinic vortex processes

Title. Author(s)OSHIMA, Hiromitsu; MAEKAWA, Tokumistu. Issue Date Doc URL. Type. File Information

Stationary Rossby Waves and Shocks on the Sverdrup Coordinate

Lecture 14. Equations of Motion Currents With Friction Sverdrup, Stommel, and Munk Solutions Remember that Ekman's solution for wind-induced transport

Convection Induced by Cooling at One Side Wall in Two-Dimensional Non-Rotating Fluid Applicability to the Deep Pacific Circulation

Long-term variability of the Kuroshio path south of Japan

Lecture #2 Planetary Wave Models. Charles McLandress (Banff Summer School 7-13 May 2005)

BALANCED FLOW: EXAMPLES (PHH lecture 3) Potential Vorticity in the real atmosphere. Potential temperature θ. Rossby Ertel potential vorticity

Driving Mechanism of Band Structure of Mean Current over the Continental Shelf

East-west SST contrast over the tropical oceans and the post El Niño western North Pacific summer monsoon

Using simplified vorticity equation,* by assumption 1 above: *Metr 430 handout on Circulation and Vorticity. Equations (4) and (5) on that handout

Bifurcation Current along the Southwest Coast of the Kii Peninsula

The Taiwan-Tsushima Warm Current System: Its Path and the Transformation of the Water Mass in the East China Sea

Warm Eddy Movements in the Eastern Japan Sea

Fixed Rossby Waves: Quasigeostrophic Explanations and Conservation of Potential Vorticity

Numerical study of the spatial distribution of the M 2 internal tide in the Pacific Ocean

Synoptic Meteorology II: Self-Development in the IPV Framework. 5-7 May 2015

LFV in a mid-latitude coupled model:

A Study on Residual Flow in the Gulf of Tongking

Measuring the Flow Through the Kerama Gap

Sensitivity of the Interannual Kuroshio Transport Variation South of Japan to Wind Dataset in OGCM Calculation

Measuring the Flow through the Kerama Gap

Cold air outbreak over the Kuroshio Extension Region

Islands in Zonal Flow*

Quasigeostrophic Dynamics of the Western Boundary Current

Measurement of Rotation. Circulation. Example. Lecture 4: Circulation and Vorticity 1/31/2017

Chapter 13 Instability on non-parallel flow Introduction and formulation

2/15/2012. Earth System Science II EES 717 Spring 2012

A Structure of the Kuroshio and Its Related Upwelling on the East China Sea Shelf Slope

Internal Wave Generation in Straits

Goals of this Chapter

The Current Structure of the Tsushima Warm Current along the Japanese Coast

Numerical Experiment on the Fortnight Variation of the Residual Current in the Ariake Sea

Marginal Sea - Open Ocean Exchange

Stability of meridionally-flowing grounded abyssal currents in the ocean

Comparison Figures from the New 22-Year Daily Eddy Dataset (January April 2015)

1/27/2010. With this method, all filed variables are separated into. from the basic state: Assumptions 1: : the basic state variables must

Eddy Formation Near the Izu-Ogasawara Ridge and its Link with Seasonal Adjustment of the Subtropical Gyre in the Pacific

Generation Mechanism of Higher Mode Nondispersive Shelf Waves by Wind Forcing

Observation of Oceanic Structure around Tosa-Bae Southeast of Shikoku

196 7 atmospheric oscillations:

Dynamics of Downwelling in an Eddy-Resolving Convective Basin

Eddy-resolving Simulation of the World Ocean Circulation by using MOM3-based OGCM Code (OFES) Optimized for the Earth Simulator

Gravity Waves. Lecture 5: Waves in Atmosphere. Waves in the Atmosphere and Oceans. Internal Gravity (Buoyancy) Waves 2/9/2017

weak mean flow R. M. Samelson

Seasonal Variations of Water Properties and the Baroclinic Flow Pattern in Toyama Bay under the Influence

The Planetary Circulation System

1/25/2010. Circulation and vorticity are the two primary

Modeling of Coastal Ocean Flow Fields

2. Baroclinic Instability and Midlatitude Dynamics

Modeling of deep currents in the Japan/East Sea

Generation and Evolution of Internal Waves in Luzon Strait

A Statistical Investigation of Internal Wave Propagation in the Northern South China Sea

Coastal-Trapped Waves with Several-Day Period Caused by Wind along the Southeast Coast of Honshu, Japan

Topographic Effects on Stratified Flows

Physical Oceanography, MSCI 3001 Oceanographic Processes, MSCI Dr. Katrin Meissner Ocean Dynamics.

Eddies, Waves, and Friction: Understanding the Mean Circulation in a Barotropic Ocean Model

Instability of a coastal jet in a two-layer model ; application to the Ushant front

2013 Annual Report for Project on Isopycnal Transport and Mixing of Tracers by Submesoscale Flows Formed at Wind-Driven Ocean Fronts

Tilting Shear Layers in Coastal Flows

Note that Rossby waves are tranverse waves, that is the particles move perpendicular to the direction of propagation. f up, down (clockwise)

Basin-Scale Topographic Waves in the Gulf of Riga*

P-Vector Inverse Method Evaluated Using the Modular Ocean Model (MOM)

Physical Oceanographic Context of Seamounts. Pierre Dutrieux Department of Oceanography, University of Hawai'i at Manoa, Honolulu, Hawai'i

A Truncated Model for Finite Amplitude Baroclinic Waves in a Channel

Final Examination, MEA 443 Fall 2008, Lackmann

Laboratory Modeling of Internal Wave Generation in Straits

Variability of Current Structure Due to Meso-Scale Eddies on the Bottom Slope Southeast of Okinawa Island

Influence of the Seasonal Thermocline on the Intrusion of Kuroshio across the Continental Shelf Northeast of Taiwan

Wind-driven Western Boundary Ocean Currents in Terran and Superterran Exoplanets

Rotating stratified turbulence in the Earth s atmosphere

A NUMERICAL EXPERIMENT OF 50-DAY RESONANCE INDUCED BY INDIAN OCEAN KELVIN WAVE IN THE SULAWESI SEA

Circulation in the South China Sea in summer of 1998

Effects of a topographic gradient on coastal current dynamics

Boundary Layers: Homogeneous Ocean Circulation

Superparameterization of Oceanic Boundary Layer Transport

Modeling of Coastal Ocean Flow Fields

Large-Scale Circulation with Locally Enhanced Vertical Mixing*

Fronts in November 1998 Storm

Anticyclonic Eddy Revealing Low Sea Surface Temperature in the Sea South of Japan: Case Study of the Eddy Observed in

NOTES AND CORRESPONDENCE. The Spindown of Bottom-Trapped Plumes

Internal boundary layers in the ocean circulation

Transcription:

Journal of Oceanography, Vol. 56, pp. 409 to 418, 2000 Numerical Study of the Generation and Propagation of Trigger Meanders of the Kuroshio South of Japan TAKAHIRO ENDOH* and TOSHIYUKI HIBIYA Department of Earth and Planetary Physics, Graduate School of Science, The University of Tokyo, Tokyo 113-0033, Japan (Received 13 October 1999; in revised form 17 December 1999; accepted 17 December 1999) We examine the processes underlying the generation and propagation of the small meander of the Kuroshio south of Japan which occurs prior to the transition from the non-large meander path to the large meander path. The study proceeds numerically by using a two-layer, flat-bottom, quasi-geostrophic inflow-outflow model which takes account of the coastal geometries of Kyushu, Nansei Islands, part of the East China Sea, and the Izu Ridge. The model successfully reproduces the observed generation and propagation features of what is called trigger meander until it passes by Cape Shiono-misaki; presumably because of the absence of the bottom topography, the applicability of the present numerical model becomes questionable after the trigger meander passes by Cape Shiono-misaki. The generation of the trigger meander off the south-eastern coast of Kyushu is shown to be associated with the increase in the supply of cyclonic vorticity by the enhanced current velocity in the upper layer along the southern coast of Kyushu where the no-slip boundary condition is employed. Thereafter, the trigger meander propagates eastward while inducing an anticyclone-cyclone pair in the lower layer. The lower-layer cyclone induced in this way, in particular, plays a crucial role in intensifying the trigger meander trough via cross-stream advection in the upper layer; the intensified trigger meander trough then further amplifies the lower-layer cyclone. This joint evolution of the upper-layer meander trough and the lower-layer cyclone indicates that baroclinic instability is the dominant mechanism underlying the rapid amplification of the eastward propagating trigger meander. Keywords: Trigger meanders, Kuroshio, numerical simulation, baroclinic instability, two-layer quasigeostrophic model, Tokara Strait, coastal geometry of Kyushu. 1. Introduction It is well known that the Kuroshio path south of Japan exhibits remarkable bimodal features, namely, the large meander (LM) path and the non-large meander (NLM) path (see Fig. 1). These bimodal features cannot be found in other western boundary currents such as the Gulf Stream. The Kuroshio path variations exert a strong influence on fisheries, ship navigation, and marine resources so that the bimodal features of the Kuroshio have attracted the attention of many researchers for a long time. One of the important observed features is that the transition from the NLM path to the LM path is preceded by the amplification of a small meander during the course of the eastward propagation after generation off the south- * Corresponding author. E-mail: enchan@aos.eps.s.utokyo.ac.jp Copyright The Oceanographic Society of Japan. eastern coast of Kyushu. The small meander is therefore often called the trigger meander (Solomon, 1978). By analyzing the time series of sea level difference between Naze and Nishinoomote (see Fig. 1) from 1965 to 1992, Kawabe (1995) showed that the Kuroshio current velocity through the Tokara Strait once increased before the trigger meander was formed, and, except for 1975, the trigger meander was generated while thus increased Kuroshio current velocity was relaxing. Although several numerical studies have been carried out to investigate the transition from the NLM path to the LM path preceded by the eastward propagation of the trigger meander (Sekine and Toba, 1981; Chao, 1984; Yasuda et al., 1985; Yoon and Yasuda, 1987; Akitomo et al., 1991), such timedependent features observed in the Tokara Strait have not been taken into account. An exception is the numerical study by Akitomo et al. (1997), but their study is limited to the barotropic responses of the Kuroshio. Considering that the time variation of sea level difference between 409

Fig. 1. Typical paths of the Kuroshio. nnlm is the nearshore non-large meander (NLM) path; onlm is the offshore NLM path; tlm is the typical large meander (LM) path. Thin solid line shows 500 m isobath (adapted from Kawabe, 1995). Naze and Nishinoomote mostly reflects that of the surface geostrophic current velocity through the Tokara Strait, the study of baroclinic responses of the Kuroshio is indispensable as well. This motivated us to carry out numerical experiments using a two-layer quasi-geostrophic (QG) model. In the present study we first demonstrate that the simple twolayer, flat-bottom, inflow-outflow model can reproduce the observed generation and propagation features of the trigger meander remarkably well, until the trigger meander passes by Cape Shiono-misaki. Next, by analyzing the calculated results, we show that lateral shear of the Kuroshio is relevant to the generation of the trigger meander off the south-eastern coast of Kyushu, whereas the effect of vertical current shear is essential in the amplification of the eastward propagating trigger meander. It should be noted that, presumably because of the absence of the bottom topography, the applicability of the present numerical model becomes questionable after the trigger meander passes by Cape Shiono-misaki so that the present study does not discuss the stationary path of the Kuroshio including the transition from the NLM path to the LM path. 2. Model The basic equations are the QG equations for a twolayer ocean given by 2 f0 Hψ 1 + t D h 1 f0 = ( ) + D J h J f 2 1 1 H A 4 ψ, ψ, ψ1 H Hψ, 1 ( )+ 1 () 1 2 f0 Hψ 2 t D h 2 f0 = ( ) + D J h J f 2 2 2 H A 4 ψ, ψ, ψ2 H Hψ, 2 ( )+ 2 ( 2) where ψ 1 and ψ 2 are the stream functions in the upper and lower layers, respectively, defined by u i i i = ψ vi = ψ, i = 1, 2 3 y x ( ) ( ) where (u i, v i ) are the corresponding horizontal current velocities; the upper and lower layer thicknesses in the undisturbed state, D 1 and D 2, are assumed to be 500 m and 3500 m, respectively; the horizontal eddy viscosity coefficient A H is assumed to be 500 m 2 s 1 ; J(A, B) = A/ x B/ y A/ y B/ x is the Jacobian operator; f is the Coriolis parameter defined by 410 T. Endoh and T. Hibiya

Fig. 2. Model geometry and coordinate system. We assume a narrow jet inside the dash-dotted line as an initial state. The shaded region is a buffer region where the large value of horizontal eddy viscosity is assumed. ( ) ( ) f = f + β xsinθ + ycos θ, 0 4 where f 0 = 7 10 5 s 1, β = 2 10 11 m 1 s 1, and taking account of the effect of the tilted coastline from Kyushu to the Izu Ridge, we rotate the (x, y) axes counter clockwise by θ = 20 degrees from the eastward and the northward directions, respectively (Chao, 1984; Yoon and Yasuda, 1987; Yamagata and Umatani, 1989; Sekine, 1990; Akitomo et al., 1991, 1997); and h is the interface displacement defined by f h = 0 ψ g ψ, ( ) ( ) 2 1 5 where the reduced gravity g* is 0.02 ms 2. The QG equations (1) and (2) are replaced with a finite difference scheme by applying the centered difference leapfrog scheme. In particular, the Arakawa Jacobian is used for an expression of the advective term (Arakawa, 1966). The model geometry is essentially the same as that presented in Akitomo et al. (1997) (see Fig. 2). This model basin is divided into square grids with a horizontal spacing of 10 km. The Kuroshio is driven in the upper layer by assuming inflow at the lower left-hand corner and outflow at the upper right-hand corner. At the outlet, the Kuroshio, with a triangular velocity profile 100 km wide, is assumed to have maximum current velocity U(t), whereas at the inlet we assume the boundary conditions given by 2 ψi = ψi = 0 i = 1, 2. 6 y y ( ) ( ) As an initial state we assume a narrow jet (just inside the dash-dotted line in Fig. 2) with the same profile as that at the outlet all the way from the inlet down to the outlet, and then we let it go. Until transient responses die away and hence quasi-stationary flow is attained, U(t) at the outlet is kept constant at U b. Thereafter we change the value of ψ 1 at the coast of Nansei Islands as well as at the right-hand side and lower side boundaries such that pulselike variation with magnitude U is superposed on U b at the outlet (Fig. 3). Associated with this, pulse-like variation of the averaged Kuroshio current velocity occurs in the upper layer in the Tokara Strait. Although the variation of the Kuroshio current velocity in the Tokara Strait, reproduced in this way, might be too simple to simulate the observed results shown in Fig. 3, the same time-dependent boundary condition was also employed in the barotropic model of Akitomo et al. (1997), which makes it easier to see the difference between the barotropic and baroclinic responses of the Kuroshio. All lateral boundaries are assumed to be no-slip except at the coast of the Izu Ridge. To prevent the computational noise generated at the inlet and the outlet from affecting the interior region, a high viscosity zone is assumed in the shaded region in Fig. 2, where the horizontal eddy viscosity coefficient A H is increased linearly from 500 up to 2500 m 2 s 1 as the right-hand side and lower side boundaries are approached. The numerical computation proceeds using a discrete time step of 3600 s. To avoid numerical instability, the Euler backward scheme (Matsuno, 1966) is applied five times successively at every 100 time steps. Numerical Study of Trigger Meanders of the Kuroshio 411

Fig. 3. The left-hand panel shows the time series of the maximum current velocity of the Kuroshio assumed at the outlet (see Fig. 2) which corresponds to the time variations of sea level difference between Naze and Nishinoomote (for the geographical locations, see Fig. 1) observed in 1986 and 1989 shown in the right-hand panel (Kawabe, 1995). The arrow in the right-hand panel shows the phase at which the trigger meander was generated. Fig. 4. Comparison of the streamline patterns in the upper layer for U b = 0.8 ms 1 and U = 0.55 ms 1 with the observed flow patterns in 1989 reported by the Japan Maritime Safety Agency (Oceanographic Prompt Reports). Contour interval for the streamlines is 5 10 3 m 2 s 1. Dashed contours are negative. 412 T. Endoh and T. Hibiya

3. Results We carried out a total of 63 numerical experiments by varying U b from 0.4 ms 1 to 1.0 ms 1 at intervals of 0.1 ms 1 and U from 0.3 ms 1 to 0.7 ms 1 at intervals of 0.05 ms 1. For U < 0.4 ms 1, no appreciable small meander can be reproduced off the south-eastern coast of Kyushu, independent of the values of U b. For U 0.4 ms 1, by contrast, the small meander generated off the south-eastern coast of Kyushu becomes prominent, although it decays during the subsequent eastward propagation unless U b 0.5 ms 1. In order to reproduce the generation and propagation features of the trigger meander, therefore, these parameters should be in the range of U b 0.5 ms 1 and U 0.4 ms 1. As U becomes larger for each U b, the amplitude of the trigger meander generated off the south-eastern coast of Kyushu becomes larger, whereas, as U b becomes larger for each U, the trigger meander thus generated amplifies more rapidly during the subsequent eastward propagation. As representative examples, Figs. 4 and 5 show the calculated results for U b = 0.8 ms 1 and U = 0.55 ms 1 and for U b = 0.6 ms 1 and U = 0.6 ms 1, which are found among others to simulate most closely the observed stream patterns for 1989 (Figs. 4a f ) and for 1986 (Figs. 5a f ) reported by the Japan Maritime Safety Agency (Oceanographic Prompt Reports). Although it is impossible to validate the chosen values of U b with the observed results, U 0.6 ms 1 is consistent with the observed result that the sea level difference between Naze and Nishinoomote increased by about 0.3 m (Fig. 3), so long as it is caused by the enhancement of the upper-layer geostrophic current velocity through the Tokara Strait. We now briefly describe the calculated results, checking by means of a comparison with the observed stream patterns. The phase for each streamline pattern is shown in Fig. 3. As U(t) increases and hence approaches the maximum, the Kuroshio gradually separates from the south-eastern coast of Kyushu (see Figs. 4a and 5a). The separation continues even while U(t) is decreasing, and by the end of the decelerating stage (t 60 days), the trigger meander is generated off the south-eastern coast of Kyushu (Figs. 4b and 5b). The trigger meander generated in this way is seen to be amplified while propagating eastward with an anticyclonic eddy behind it (Figs. 4c f and 5c f ). Although the agreement between the calculated results and the observed stream patterns for 1986 Fig. 5. Comparison of the streamline patterns in the upper layer for U b = 0.6 ms 1 and U = 0.6 ms 1 with the observed flow patterns in 1986 reported by the Japan Maritime Safety Agency (Oceanographic Prompt Reports). Contour interval for the streamlines is 5 10 3 m 2 s 1. Numerical Study of Trigger Meanders of the Kuroshio 413

is not as good as that for 1989, the general characteristics of the observed features are reproduced well as a whole, implying that the essential physics of the phenomenon is successfully simulated in the present simple two-layer model. In the barotropic model of Akitomo et al. (1997), the trigger meander was generated dozens of days after the short-term variation of the Kuroshio current velocity was applied in the Tokara Strait, which is inconsistent with the observed results. Considering that the model geometry and initial and time-dependent boundary conditions in Akitomo et al. (1997) are essentially the same as those in the present study, this inconsistency can be attributed to the neglect of baroclinicity. 4. Discussion Figure 6 shows the areas of relative vorticity exceeding 1 10 5 s 1 superposed on the calculated streamline pattern at t = 60 days for the representative case shown in Fig. 4. A narrow band of cyclonic vorticity is seen to extend from the southern coast of Kyushu to the trough of the trigger meander, suggesting that the generation of the trigger meander is associated with the relative vorticity supplied at the southern coast of Kyushu where the noslip boundary condition is employed. This is more directly confirmed by the fact that the generation of the trigger meander does not occur at all in the numerical experiment with all the lateral boundaries assumed to be freeslip (results not shown). Figure 7 shows the time variation of the maximum current velocity in the upper layer through the Tokara Strait, whereas Fig. 8 provides that of the term balances in Eq. (1) at the location (x, y) = (350 km, 210 km) marked by a star ( ) in Fig. 6. After the current velocity in the upper layer along the southern coast of Kyushu attains the maximum at t 40 days, the relative vorticity at this location rapidly increases, mainly through the advection of cyclonic vorticity from the Tokara Strait. The enhancement of the Kuroshio along Fig. 7. Time variation of the maximum current velocity in the upper layer through the Tokara Strait for the case shown in Fig. 4. Fig. 6. Spatial distribution of cyclonic vorticity H 2 ψ 1 exceeding 1 10 5 s 1 (shaded regions) superposed on the ψ 1 fields (thin lines) at t = 60 days in Fig. 4. A star ( ) indicates the location where the time variation of the term balances in the vorticity equation (1) are examined (see Fig. 8). Fig. 8. Time variation of the term balances in the vorticity equation (1) at the location marked by in Fig. 6 for the case shown in Fig. 4;, advection of relative vorticity J(ψ 1, H 2 ψ 1 );, horizontal dissipation A H H 4 ψ 1 ;, advection of planetary vorticity J(ψ 1, f );, vertical stretching f 0 (( h/ t) + J(ψ 1, h))/d 1 ;, local time change of relative vorticity ( / t)( H 2 ψ 1 ). 414 T. Endoh and T. Hibiya

Fig. 9. Calculated streamline patterns in the upper layer using the reduced gravity model where the same values of U b and U are assumed as those in Fig. 4. Contour interval is 5 10 3 m 2 s 1. the southern coast of Kyushu thus increases the viscous production as well as downstream advection of cyclonic vorticity, resulting in the generation of the trigger meander off the south-eastern coast of Kyushu. This suggests that the same generation process might be reproduced using a reduced gravity model. In fact, the reduced gravity model with the same values of U b and U as assumed in Fig. 4 reproduces the small meander of nearly the same amplitude at t 60 days (Fig. 9). In contrast, the propagation features of the small meander are quite different between the two-layer model and the reduced gravity model. Figure 10 shows the time variations of the total amplitude of the small meanders shown in Figs. 4 and 9, respectively. The trigger meander in the two-layer model rapidly amplifies during t = 90 120 days, quite different from the small meander in the reduced gravity model, which decays during the same time period. To understand the underlying mechanism for the rapid amplification of the trigger meander reproduced in the two-layer model, we focus on the first term on the right-hand side of Eq. (1), namely, f 0 J(ψ 1, h)/d 1 which vanishes in a reduced gravity model. Figure 11(a) shows the contours of f 0 J(ψ 1, h)/d 1 (thick lines) superposed on Fig. 10. Time variations of the total amplitude of the small meanders shown in Fig. 4 ( ) and in Fig. 9 ( ). Numerical Study of Trigger Meanders of the Kuroshio 415

the ψ 1 fields (thin lines) for the case in Fig. 4. Until the rapid amplification of the meander trough almost terminates at t 110 days (see Fig. 10), large negative values of f 0 J(ψ 1, h)/d 1 can be found at the location of the steepening trough. From Eq. (1) we can see that the negative values of f 0 J(ψ 1, h)/d 1 correspond to the cross-stream advection, causing partial relaxation of the interface displacement as well as vertical stretching of upper-layer parcels, accompanied by enhancement of cyclonic vorticity (Cushman-Roisin, 1994). It is interesting to note that, using Eq. (5), f 0 J(ψ 1, h)/d 1 can alternatively be written as f 0 2 J(ψ 1, ψ 2 )/g*/d 1 so that the strength of the crossstream advection effect in the upper layer is greatest when the induced lower-layer currents flow at large angles to the upper-layer Kuroshio. In fact, Fig. 11(b) shows that the trigger meander amplifies rapidly during t = 90 120 days when an anticyclone-cyclone pair below the meander trough-crest pair evolves, but the trigger meander ceases to amplify after t 120 days when the angle between the upper- and lower-layer currents becomes fairly small and, furthermore, the cyclone below the steepened meander trough decays, presumably through the lateral friction at the southern coast of Japan. The joint evolution of the upper-layer meander trough and the lowerlayer cyclone is also confirmed to occur for the case shown in Fig. 5 (results not shown), indicating that baroclinic instability is the important mechanism underlying the rapid amplification of the eastward propagating trigger meander. The importance of f 0 J(ψ 1, h)/d 1 was first pointed out by Yoon and Yasuda (1987), who demonstrated the relevance of baroclinic instability to the amplification of the trigger meander, based on the idealized two-layer numerical model. Even after their study, however, the dynamics of transient responses of the Kuroshio and the stationary path of the Kuroshio following the transient processes have been discussed using the barotropic numerical model (e.g. Yamagata and Umatani, 1989; Akitomo et al., 1991, 1997). The main reason for this is that there has been no attempt to check the validity of each model result through comparison with the observed stream patterns. The numerical model of Yoon and Yasuda (1987), for example, was somewhat too simple to allow comparison with the observed stream patterns where the trigger meander was introduced in the upper layer a priori in the form of a cyclonic eddy, with the generation process being obscured. The emphasis of the present study is placed on the demonstration that the observed generation and propagation features of the trigger meander can be simulated very well using the two-layer, quasi-geostrophic model, which includes the coastal geometries of Kyushu, Nansei Islands, part of the East China Sea, and Izu Ridge, and, more specifically, the mechanism of baroclinic instability. 5. Concluding Remarks Using a two-layer, flat-bottom, quasi-geostrophic numerical model, we have reproduced the generation and propagation features of the trigger meander of the Kuroshio which agree remarkably well with the observed ones until the trigger meander passes by Cape Shionomisaki. The generation of the trigger meander off the south-eastern coast of Kyushu is associated with the increase in the supply of cyclonic vorticity caused by the enhanced current velocity in the upper layer along the southern coast of Kyushu where the no-slip boundary condition is employed. Thereafter, the trigger meander propagates eastward while inducing an anticyclone-cyclone pair in the lower layer. The lower-layer cyclone induced in this way, in particular, plays a crucial role in intensifying the trigger meander trough via cross-stream advection in the upper layer; the intensified trigger meander trough then further amplifies the lower-layer cyclone. This joint evolution of the upper-layer meander trough and the lower-layer cyclone indicates that baroclinic instability is the dominant mechanism underlying the rapid amplification of the eastward propagating trigger meander, consistent with the results of Yoon and Yasuda (1987) based on the idealized two-layer numerical model. Several important problems still remain to be investigated. First, the effect of bottom topography should be clarified. When a realistic bottom topography is taken into account, the Kuroshio tends to flow along a continental shelf slope instead of flowing into the offshore, deeper, flat-bottom region where it is susceptible to baroclinic instability (Sekine, 1992). To clarify the effect of bottom topography, we have to take a more precise account of the existence of density stratification. Ikeda (1983), for example, showed that at least a three-layer model should be used to study the instability of a current flowing along a continental shelf slope. The absence of the bottom topography in the present numerical model might also be responsible for the disagreement between the calculated stream pattern and the observed one after the trigger meander passes by Cape Shiono-misaki. For example, we could not reproduce what is called the S-shaped path where the Kuroshio loops back west of the Izu Ridge just before the LM path is formed (Sakaida et al., 1998). Since the reliability of the stationary path of the Kuroshio that follows this transient process is considered to be questionable, the present study has not discussed the stationary path of the Kuroshio including the transition from the NLM path to the LM path. Second, it is not clear what causes the short-term variation of the Kuroshio current velocity in the Tokara Strait, although the interaction between the Kuroshio and a mesoscale eddy approaching the Tokara Strait might be a plausible explanation. To answer these questions, we are planning a detailed nu- 416 T. Endoh and T. Hibiya

Fig. 11. (a) Contours of f 0 J(ψ 1, h)/d 1 (thick lines) superposed on the ψ 1 fields (thin lines) for the case shown in Fig. 4. Contour intervals are 5 10 12 s 2 for f 0 J(ψ 1, h)/d 1 and 1 10 4 m 2 s 1 for ψ 1. The areas where f 0 J(ψ 1, h)/d 1 5 10 12 s 2 are shaded. (b) The ψ 2 fields (thick lines) superposed on the ψ 1 fields (thin lines) for the case shown in Fig. 4. Contour intervals are 2 10 3 m 2 s 1 for ψ 2 and 1 10 4 m 2 s 1 for ψ 1. Dashed contours are negative. Numerical Study of Trigger Meanders of the Kuroshio 417

merical study based on a 3-D, primitive, high resolution level model that takes account of realistic topography, the results of which will be reported elsewhere. Acknowledgements The authors express their gratitude to Dr. M. Kawabe of the Ocean Research Institute for fruitful discussions. They are also grateful to Prof. T. Matsuno of the Nagasaki University, Prof. M. Wimbush of the University of Rhode Island, and two anonymous reviewers for their invaluable comments and suggestions on the original manuscript. References Akitomo, K., T. Awaji and N. Imasato (1991): Kuroshio path variation south of Japan: 1. Barotropic inflow-outflow model. J. Geophys. Res., 96, 2549 2560. Akitomo, K., S. Masuda and T. Awaji (1997): Kuroshio path variation south of Japan: Stability of the paths in a multiple equilibrium regime. J. Oceanogr., 53, 129 142. Arakawa, A. (1966): Computational design for long-term numerical integration of the equation of fluid motion: Twodimensional incompressible flow, Part 1. J. Comput. Phys., 1, 119 143. Chao, S.-Y. (1984): Bimodality of the Kuroshio. J. Phys. Oceanogr., 14, 92 103. Cushman-Roisin, B. (1994): Introduction to Geophysical Fluid Dynamics. Prentice Hall, New Jersey, 320 pp. Ikeda, M. (1983): Linear instability of a current flowing along a bottom slope using a three-layer model. J. Phys. Oceanogr., 13, 208 223. Kawabe, M. (1995): Variations of current path, velocity, and volume transport of the Kuroshio in relation with the large meander. J. Phys. Oceanogr., 25, 3103 3117. Matsuno, T. (1966): Numerical integration of the primitive equations by a simulated backward difference method. J. Meteorol. Soc. Japan, 44, 76 84. Sakaida, F., D. Egusa and H. Kawamura (1998): The behavior and the role of the anti-cyclonic eddy in the Kuroshio large meander development. J. Oceanogr., 54, 101 114. Sekine, Y. (1990): A numerical experiment on the path dynamics of the Kuroshio with reference to the formation of the large meander path south of Japan. Deep-Sea Res., 37, 359 380. Sekine, Y. (1992): On the signs of formation of the Kuroshio large meander south of Japan. Bull. Japan. Soc. Fish. Oceanogr., 56, 13 22 (in Japanese). Sekine, Y. and Y. Toba (1981): A numerical study on the generation of the small meander of the Kuroshio off southern Kyushu. J. Oceanogr. Soc. Japan, 37, 234 242. Solomon, H. (1978): Occurrence of small trigger meanders in the Kuroshio off southern Kyushu. J. Oceanogr. Soc. Japan, 34, 81-84. Yamagata, T. and S. Umatani (1989): Geometry-forced coherent structures as a model of the Kuroshio large meander. J. Phys. Oceanogr., 19, 130 138. Yasuda, I., J.-H. Yoon and N. Suginohara (1985): Dynamics of the Kuroshio large meander: Barotropic model. J. Oceanogr. Soc. Japan, 41, 259 273. Yoon, J.-H. and I. Yasuda (1987): Dynamics of the Kuroshio large meander: Two-layer model. J. Phys. Oceanogr., 17, 66-81. 418 T. Endoh and T. Hibiya