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

Size: px
Start display at page:

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

Transcription

1 994 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32 Annual Variation of the Kuroshio Transport in a Two-Layer Numerical Model with a Ridge ATSUHIKO ISOBE Department of Earth System Science and Technology, Interdisciplinary Graduate School of Engineering Sciences, Kyushu University, Kasuga, Fukuoka, Japan SHIRO IMAWAKI Research Institute for Applied Mechanics, Kyushu University, Kasuga, Fukuoka, Japan (Manuscript received 28 September 2000, in final form 3 August 2001) ABSTRACT A two-layer numerical model driven by the wind stress is used to explain the observed annual variation of the Kuroshio transport south of Japan. Special attention is given to the effect of a ridge, representing the Izu Ogasawara Ridge, on the generation of the baroclinic activity through the coupling of the barotropic and baroclinic modes of motion. For annual variation, lower-layer motion is found in areas surrounding the ridge because isostasy (a state of no motion in the lower layer) is not achieved within such a short timescale. Thus, the lowerlayer flow impinges on the bottom slope. This impinging process generates anomalies of the upper-layer thickness especially on the eastern side of the ridge. Thereafter, anomalies move westward with characteristic velocities composed of the vertically averaged flow and westward propagation of the long baroclinic Rossby wave forced above the ridge. As anomalies of the upper-layer thickness move westward above the ridge, isostasy is accomplished with respect to these anomalies. As a result, the positive (negative) anomaly of the upper-layer thickness carries the information about the positive (negative) anomaly of the volume transport as it reaches the western edge of the ridge. Thereafter, anomalies of the volume transport are released to the west of the ridge. This experiment shows that the annual range of the volume transport east of the ridge is around 40 Sv, which is nearly equal to the zonal integration of the Sverdrup transport there. The annual range west of the ridge, however, reduces to around 10 Sv, which is mostly caused by the baroclinic activity generated above the ridge. Results are compared with the observed Kuroshio transport across the ASUKA line south of Japan. The annual range west of the ridge is consistent with that estimated from the observation. 1. Introduction The annual cycle is the most energetic component of the wind stress driving the general ocean circulation. Therefore, it is reasonable to anticipate a prominent annual cycle in the Sverdrup transport. Moreover, it is anticipated that the zonal integration of the Sverdrup transport well describes the annual transport variation of the western boundary current as the compensation flow of the interior transport, because the barotropic adjustment is achieved instantaneously compared to the annual timescale. In the case of the Gulf Stream, however, the annual range of the volume transport variation is less than 10 Sv (1 Sv 10 6 m 3 s 1 ) in the Florida Strait (Niiler and Richardson 1973), which is much smaller than that of the Sverdrup transport there ( 30 Corresponding author address: Dr. Atsuhiko Isobe, Department of Earth System Science and Technology, Interdisciplinary Graduate School of Engineering Sciences, Kyushu University, 6-1, Kasuga- Koen, Kasuga, Fukuoka , Japan. isobe@esst.kyushu-u.ac.jp Sv; Anderson and Corry 1985). Such an annual cycle of the Gulf Stream has been confirmed by Molinari et al. (1985) and Schott and Zantopp (1985) for the Florida Strait, and by Sato and Rossby (1995) for the east of Cape Hatteras, although the transport variation due to interactions between the stream and surrounding eddy field makes it difficult to determine an accurate annual cycle. In the case of the Kuroshio, the western boundary current of the North Pacific subtropical gyre, recent observations by the Affiliated Surveys of the Kuroshio off Cape Ashizuri (ASUKA) Group provided a time series of volume transport for seven years (Imawaki et al. 1997, 2001; H. Uchida and S. Imawaki 2001, personal communication, hereafter UI). They carried out moored current-meter measurements and repeated hydrographic surveys along a line shown in Fig. 1 (ASUKA line), and estimated the absolute geostrophic transport of the Kuroshio across the ASUKA line. In addition, they found an empirical formula converting the altimetry data to the volume transport of the Kuroshio. As a result, 2002 American Meteorological Society

2 MARCH 2002 ISOBE AND IMAWAKI 995 FIG. 1. Study area of the western North Pacific and results of the ASUKA observations. Thick solid lines represent the ASUKA and PN lines. A thick broken line shows the path of the Kuroshio schematically. Also shown are isobaths in kilometers. The solid curve in the lower-right panel shows the annual variation of the Kuroshio volume transport (after UI. The broken line is the annual variation of the negative value of the Sverdrup transport integrated over the North Pacific at 30 N, evaluated from ECMWF wind data (K. Kutsuwada 1999, personal communication). they are now monitoring the Kuroshio volume transport every 10 days using the satellite altimetry data. According to the climatological mean of the geopotential anomaly field depicted by Hasunuma and Yoshida (1978), a stable anticyclonic eddy exists south of Shikoku Island. This eddy might prevent an accurate estimate of the Kuroshio volume transport. However, the ASUKA line from Cape Ashizuri to 25 N extends to the south beyond the local eddy. Hence, the contribution of the eddy is cancelled effectively in their estimate of the Kuroshio volume transport. This means that the annual cycle in the observed Kuroshio volume transport originates from the annual variability of gyre-scale motion. Nevertheless, as shown in Fig. 1, the averaged annual range of the volume transport variation is estimated to be about 10 Sv, which is much smaller than that of zonally integrated Sverdrup transport (40 50 Sv) at 30 N over the North Pacific (K. Kutsuwada 1999, personal communication); this transport is estimated from wind data of the European Center for Medium- Range Weather Forecasts (ECMWF) with the drag coefficient of Large and Pond (1981). The actual ocean has a complicated bottom topography, so the classical (i.e., nontopographic) Sverdrup relation is only applicable to phenomena with a timescale longer than the time for baroclinic Rossby waves to cross the ocean and compensate for the effect of the bottom topography (Anderson and Killworth 1977). For the midlatitude Atlantic and Pacific, it takes years or decades for compensation by baroclinic Rossby waves to occur. Therefore, for shorter period (annual or less), the stratification is unlikely to filter out the effect of topography. Thus, one may consider that the volume transport variation for shorter period is explained by the Sverdrup relation for a homogeneous ocean with topography, that is, the topographic Sverdrup relation described by Gill and Niiler [1973, see their Eq. (5.9) and discussion] and Thompson et al. (1986). This means

3 996 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32 FIG. 2. Two-layer model adopted in this study. The upper panel shows the plane view of the model ocean. Thin lines in the panel represent the grid spacing; it is shown only in the central part of the domain for the y direction. Meridional distribution of the wind stress is depicted schematically on the right-hand side of this panel. The lower panel shows the vertical view of the model. A ridge and the shelf slope are depicted in the panel. that a one-layer numerical model (i.e., barotropic model) with realistic topography can reproduce the annual transport variation. In fact, the phase of the annual transport variation is well reproduced by the one-layer models by Anderson and Corry (1985) and Greatbatch and Goulding (1989) for the Gulf Stream, and by Greatbatch and Goulding (1990) and Kubota et al. (1995) for the Kuroshio. However, they all underestimate the annual ranges compared to the observed variation. Namely, the observed annual ranges of the volume transport of the Gulf Stream and Kuroshio are smaller than those of the nontopographic Sverdrup transport, but larger than those of the topographic Sverdrup transport. Anderson and Corry (1985) adopted a two-layer numerical model with realistic topography as well as the one-layer model, and found that the annual range of the volume transport variation of the Gulf Stream is amplified by baroclinic activity associated with a baroclinic Kelvin wave. The Kelvin wave is generated by winds to the north of the Florida Straits, passing south over varying bottom topography. However, they still underestimate the annual range. In the present study, we also use a two-layer numerical model to identify the baroclinic activity inducing the observed annual range of the Kuroshio volume transport. As shown later, the annual range in the present model is consistent with the observed value on the ASUKA line. As shown in Fig. 1, the ASUKA line is located west of the Izu Ogasawara Ridge which extends meridionally over the whole region of the North Pacific subtropical gyre. In the present study, special attention is given to the effect of the ridge on the generation of a baroclinic Rossby wave through the coupling of the barotropic and baroclinic modes of motion (e.g., Barnier 1988; Gerdes and Wübber 1991). As shown later, the annual range of the modeled Kuroshio volume transport variation is larger than that of the topographic Sverdrup transport because of baroclinic motion excited above the ridge. The two-layer numerical model in the present study is described in section 2. In section 3, we show results in which the baroclinic activity above the ridge controls the annual variation of the Kuroshio volume transport. A physical interpretation and the comparison with the ASUKA observation are provided in section 4. Section 5 gives the conclusions. 2. Model descriptions To provide a simple interpretation for the annual variation of the Kuroshio volume transport, a rectangular ocean ( km 2000 km) with a simplified bottom topography is adopted in this model. We use Cartesian coordinates where directions of x and y axes (with origin at the southwestern corner) are eastward and northward, respectively. As shown in Fig. 2, the model ocean is flat except for a meridional ridge in the western part (1550 km x 2050 km) and the shelf slope at the western boundary. The ridge represents the Izu Ogasawara Ridge in the western North Pacific (see appendix A for its applicability to the real ocean). In the present study, the depth H of the ridge is modeled as Hr 2 H(l) H0 1 sin l, 2 W 2 where H 0 is the depth in the flat-bottom region, H r the maximum height of the ridge, W the width of the ridge, and l the eastward distance from the western edge of

4 MARCH 2002 ISOBE AND IMAWAKI 997 TABLE 1. Parameters used in the present study. Reduced gravity Planetary Coriolis parameter at southern sidewall Horizontal viscosity coefficient Amplitude of the wind stress Depth in the flat-bottom region Maximum height of the ridge Width of the ridge Undisturbed upper-layer thickness g : 2.0 cm s 2 : cm 1 s 1 f (y 0): s 1 A h : cm 2 s 1 a: dyn cm 2 H 0 : 3000 m H r : 1500 m W: 500 km H 1 : 600 m the ridge. In order to avoid computational instability, a shelf slope of 100 km width is imposed along the western boundary, prohibiting the width of the western boundary layer from being too thin. The model ocean is divided into two active layers. In the present study, the Kuroshio transport is defined as the volume transport between the western boundary and western edge of the ridge. This modeled transport corresponds to the observed Kuroshio transport across the whole ASUKA line (i.e., the transport of the Kuroshio Throughflow in Imawaki et al. 2001; UI). The governing equations are the same as those in Sekine and Kutsuwada (1994) except that momentum advection is not included in the present model. We neglect momentum advection because our attention is not focused on the current structure in the western boundary region where momentum advection cannot be neglected. More attention is given to processes around the ridge in the interior region far from the western boundary. Nonlinearity remains in the continuity equation because a large displacement of the interface is expected. Friction at both the bottom and layer interface is neglected. The lateral boundaries are nonslip sidewalls, as well as prohibiting the flow across them. The ridge intersects the northern and southern sidewalls, so sidewall boundary layers are produced artificially at both ends of the ridge. In addition, discontinuities in the f/h contours at both ends of the ridge might be artificial sources of vorticity (Barnier 1988). However, processes of importance to the present study occur apart from sidewalls, as shown in section 4. The governing equations with hydrostatic, rigid-lid, and Boussinesq approximations are f h2 J, g J, h1 t H H y H 2 Ah, (1) û h1 2 f ˆ g Ah û, (2) t x h 1 ˆ h 1 2 fû g Ah ˆ, t y (3) h1 u1h1 1h1 0, t x y (4) where is the vorticity of the vertically averaged flow, the (volume transport) streamfunction, f the Coriolis parameter on the plane, g the reduced gravity, h 1 and h 2 the layer thickness of the upper and lower layers respectively, the wind stress, û( u 1 u 2 ) and ˆ ( 1 2 ) the difference of the upper-layer velocity (u 1, 1 ) and lower-layer velocity (u 2, 2 )inx and y directions respectively, and J(A, B) the Jacobian ( A x B y A y B x ); otherwise notation is standard. The vorticity and streamfunction are defined as 1, (5) H uh 1 1 uh 2 2 y. (6) h h x The model is driven only by the zonal wind stress. The amplitude of the wind stress varies sinusoidally in the y direction. The wind stress is composed of mean and time-dependent parts as follows: a 2 cos y a cos t, (7) L 2 T where T is the period of the fluctuating wind stress ( 360 days in the present study), and L the total length of the model domain in the y direction. In this model, the wind stress is maximum on day 0 and day 360 (defined as midwinter), and the minimum on day 180 (midsummer). The parameter a in (7) is a constant determining the magnitude of wind forcing; its value is chosen so that the annual mean and range of wind stress curl integrated zonally take the same values as those estimated from ECMWF wind data [about 40 Sv for the annual mean and Sv for the annual range (Sv 10 6 m 3 s 1 )]. Parameters used in the present study are listed in Table 1. Equations (1) (4) are numerically integrated using the leapfrog scheme with the Matsuno scheme every 25 time steps. Equation (5) is solved using the successive over-relaxation method. The grid spacing in the x direction varies zonally (see Fig. 2); it is 20 km wide above the ridge and shelf slope at the western boundary, km wide west of the ridge, and km wide east of the ridge. The grid spacing in the y direction is fixed at 20 km. Initially, distributions of the streamfunction and upper-layer thickness are determined using linearized solutions of a flat-bottom case in the steady state, x f g H y x 1 (x, y) dx, (8) y x e h (x, y) h dx, (9) 1 0 x e 1

5 998 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32 FIG. 3. The streamfunction (upper panel) and upper-layer thickness (lower panel) in the twolayer model with the ridge, driven by the annually averaged wind stress. Broken lines represent the location of eastern and western edges of the ridge. Contour intervals for the streamfunction and upper-layer thickness are 5 Sv and 50 m. Contours within the western boundary layer are omitted because of over crowding. where the location x e denotes the eastern sidewall, H 1 is the undisturbed upper-layer thickness, and h 0 is a constant such that the spatially averaged upper-layer thickness is equal to the undisturbed value H 1. The wind stress is the annually averaged value in above equations. Time integration is continued until the end of the tenth year when the year-to-year variability disappears. Results taken from the final tenth year are discussed in the following sections. 3. Results a. Annual mean Before looking at the annual variation, the effect of the ridge on the annually averaged field is described. As expected, distributions of the streamfunction and upper-layer thickness represent the wind-driven ocean circulation in a textbook way when we use a two-layer model without a ridge, driven by the annual mean wind stress. In addition, as shown in Fig. 3, the solutions do not change, even though we use a two-layer model with the ridge, driven by the annual mean wind stress. This is because lower-layer motion vanishes in the annual mean case. We here refer to a state of no motion in the lower layer as isostasy. The achievement of isostasy in the steady state can be explained by the steady-state vorticity equation in the lower layer (Rooth et al. 1978; Rhines and Young 1982): f 2 f 2 0. (10) y h x x h y 2 2 Here viscous terms are neglected so that the above equation is valid in the Sverdrup interior region. Equation (10) means that the streamfunction of the lower layer, 2, is advected by characteristic velocities whose streamfunction is the geostrophic contour, f/h 2. Figure 4 shows the horizontal distribution of f/h 2 of the solution in Fig. 3. In the flat-bottom region east of the ridge, all geostrophic contours originate from the eastern sidewall along which the lower-layer streamfunction is defined to be zero. Hence, isostasy is achieved east of the ridge. The geostrophic contours from the eastern sidewall turn southward in the eastern half of the ridge due to large topographic on the ridge (see lower-left panel of Fig. 4 for the geostrophic contour f/h 2 and lowerright panel for f/h). Other geostrophic contours on the ridge originate from the northern and southern sidewalls of the ridge. In addition, the contours in the flat-bottom region west of the ridge originate from the southern sidewall of the ridge. Therefore, isostasy on the ridge and in the flat-bottom region west of the ridge results from the condition of no motion required along the sidewalls on the ridge. Two areas of closed geostrophic contours are revealed in the northwestern corner of the flat-bottom region west of the ridge and on the northern part of the ridge. In these areas, lower-layer motion is permitted even in the steady state because the areas are isolated from sidewalls. However, the present model does not directly force motion of the lower layer [e.g., by interfacial drag as in Rhines and Young, (1982) and Jarvis and Veronis (1994)]. Therefore, in the steady state, lower-layer isos-

6 MARCH 2002 ISOBE AND IMAWAKI 999 FIG. 4. Geostrophic contours of the lower layer ( f/h 2 ) over the flat bottom (upper panel) and those over the ridge (lower-left panel) in the case of the flow field in Fig. 3. Also shown is the horizontal distribution of contour lines of f/h (lower-right panel) over the ridge. Unit is in cm 1 s 1. Contour intervals are cm 1 s 1 for f/h 2 and cm 1 s 1 for f/h. tasy is accomplished through the whole domain of the present model. In the above experiment, the modeled ocean is driven by the annual mean wind stress. We here adopt the twolayer numerical model with the ridge, driven by the fluctuating wind stress expressed by Eq. (7). Annually averaged distributions (not shown) of the streamfunction and upper-layer thickness are identical to those in Fig. 3. It is interpreted that the steady-state solution accomplishing isostasy is superimposed on the timedependent part of the solution. Thus, the ridge has almost no effect on the annually averaged circulation. b. Annual variation Figure 5 shows the streamfunction every 60 days in the experiment with the ridge and wind stress varying in time. The volume transport has a large variation east of the ridge with an annual range of more than 40 Sv [see panels of day 180 (midsummer) and day 360 (midwinter)]. However, west of the ridge, a gyre with the volume transport of more than 40 Sv is maintained throughout the year. Figure 6 shows the time series of the streamfunction at the eastern and western edges of the ridge. Positive streamfunction represents net northward volume transport between the western boundary and each point. Also shown is the negative value of the nontopographic Sverdrup transport zonally integrated between the eastern sidewall and each point. This value is estimated by Eq. (8) using the time-varying wind stress. The annual means of the calculated volume transport on the eastern and western sides of the ridge are 40 Sv and 42 Sv, respectively. Annual ranges on the two sides are 45 Sv and 13 Sv, respectively. The annual range of the calculated volume transport is nearly the same as that of the nontopographic Sverdrup transport

7 1000 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32 FIG. 5. Horizontal distributions of the streamfunction every 60 days in the two-layer model with the ridge, driven by the wind stress varying annually. Broken lines represent the location of eastern and western edges of the ridge. Contour interval is 5 Sv. The streamfunction within the western boundary layer is omitted because of over crowding.

8 MARCH 2002 ISOBE AND IMAWAKI 1001 FIG. 6. Time series of the streamfunction (solid line) at eastern and western edges of the ridge, whose locations are shown in the upper panel. The broken line shows the time series of the nontopographic Sverdrup transport at corresponding point. east of the ridge, because the flat-bottom region extends east of the ridge. In addition, the phase of the calculated volume transport variation is nearly the same as that of the Sverdrup transport because the barotropic adjustment is achieved rapidly. However, the annual range decreases drastically west of the ridge in comparison with that of the nontopographic Sverdrup transport at the same point. On the other hand, the annual mean is nearly the same as that of the nontopographic Sverdrup transport. The cause of the reduction of the annual range west of the ridge is readily understood, considering the steady-state solution mentioned previously. The steadystate solution is equal to the annually averaged part of the solution in the time-dependent experiment. Isostasy is achieved in the steady-state solution, so that the effect of the ridge disappears. Therefore, the steady-state solution is permitted in the whole domain including west of the ridge. In contrast, isostasy is not achieved in the time-dependent part of the solution because the annual period is too short for the long baroclinic Rossby wave to cross the ocean. Thus, lower-layer motion remains only in the time-dependent part of the solution. Motion propagates westward because of the planetary effect, but the signal of the time-dependent part of the volume transport is prohibited from crossing the ridge due to the topographic effect. As a result, the annually averaged part of the solution (steady-state solution) is dominant in the western portion of the ridge. Although the annual range of the volume transport variation west of the ridge is much smaller than that of the nontopographic Sverdrup transport there, the volume transport varies in time slightly. Figure 7 shows the time series of the streamfunction at the western edge of the ridge, which is compared with the time series of the volume transport calculated by a one-layer numerical model with the ridge, driven by the same time-varying wind stress. As mentioned previously, the result in the one-layer model is described by the topographic Sverdrup relation. Above the ridge imposed in this model, almost all f/h contours entering from the east of the ridge turn to the south and encounter the southern sidewall (see the lower-right panel of Fig. 4). Thus, in the one-layer model, variability with an annual range west of the ridge is restricted to about 3 Sv. On the other hand, the annual range of the volume transport west of the ridge is about 10 Sv in the two-layer model. That is, the annual range of the volume transport in the twolayer model is smaller than the range of the nontopographic Sverdrup transport, but larger than that of the topographic Sverdrup transport. This means that the baroclinic activity amplifies the annual range of the volume transport. In the flat-bottom ocean without bottom friction, the baroclinic activity does not affect the volume transport but affects the vertical structure of the current. Therefore, the amplification of the volume transport variability in the two-layer model, as shown in Fig. 7, is

9 1002 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32 FIG. 7. Time series of the streamfunction (solid line) at the western edge of the ridge. Note that the scale of the ordinate is enlarged compared to that in Fig. 6. Broken line indicates the time series of the streamfunction at the same point in the one-layer model; the annual mean transport of the one-layer model (3 Sv) is shifted to that of the two-layer model (42 Sv). induced by the baroclinic activity generated above the ridge. As shown in Fig. 7, the phase of the volume transport in the one-layer model is nearly the same as that of the nontopographic Sverdrup transport at the same point (see lower-left panel of Fig. 6). The maximum volume transport appears on day 0 or day 360 (midwinter), while the minimum value appears on day 180 (midsummer). However, the phase in the two-layer model is delayed about one month. Figure 8 shows the y t (distance from the southern sidewall time) plot of the streamfunction along the western edge of the ridge. Thick broken lines represent times when the maximum and minimum values appear at each latitude. The annual variation is remarkable between 500 and 1500 km from the southern sidewall, so these two lines are depicted within this area. It is found that the phase is delayed northward. In the one-layer model, the phase delay is minimal because barotropic adjustment is achieved instantaneously compared to the annual timescale at all latitudes. Thus, the phase delay in the two-layer model is due to the baroclinic activity generated on the ridge. The baroclinic activity, which determines the annual range and phase of the volume transport west of the ridge, is discussed in detail in the next section. FIG. 8. Temporal variation of the streamfunction along the western edge of the ridge. The ordinate represents the northward distance (y) from the southern sidewall. Two broken lines indicate the times at which the maximum and the minimum values appear at each latitude. Contour interval is 5 Sv. 4. Discussion a. Anomaly fields We look at anomaly fields from annually averaged values because the baroclinic activity affects only the time-dependent part of the solution. Figure 9 displays anomalies of the upper-layer thickness (left panels) and streamfunction (right panels) every 60 days. The anomaly fields in the western part (0 x 3200 km) of the model domain are shown in order to emphasize the process around the ridge. As shown in the anomaly field of the upper-layer thickness on day 360, the negative anomaly is generated in the southeastern part of the ridge (labeled a ). This anomaly moves westward (see a on day 60, 120, and 180). Thereafter, the negative anomaly leaves the ridge and enters the flat-bottom region west of the ridge. Similarly, the negative anomaly of the upper-layer thickness is generated on day 180 in the northeastern part of the ridge (labeled b ). Although this anomaly also moves westward (see b on day 240, 300, and 360), its speed seems to be slow compared to that of a. In addition, an anomaly generated in the north vanishes on the western side of the ridge. It is also found that positive anomalies are generated on the eastern side of the ridge and move westward. The anomaly fields of the streamfunction show that FIG. 9. Horizontal distributions of the anomaly of the upper-layer thickness (left) and streamfunction (right) displayed every 60 days. Distribution only in the western part of the domain (0 x 3200 km) is depicted. Two long-broken lines in the central part of the domain indicate the location of the ridge. Contour intervals for the left and right panels are 20 m and 5 Sv, respectively. The short-broken lines are used for the negative anomalies. See the text for the explanations of letters a and b.

10 MARCH 2002 ISOBE AND IMAWAKI 1003

11 1004 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32 FIG. 10. Horizontal distributions of the magnitude of the impinging term above the ridge displayed every 60 days. Positive (negative) values indicate downwelling (upwelling). Contour interval is cm s 1 as shown in the lower portion of these panels. A broken line is used for the negative value. Contours near the sidewalls are omitted. an annual variation with a range of 40 Sv appears east of the ridge. On the contrary, the annual range of the volume transport is small west of the ridge. It is interesting to note that the negative (positive) anomaly of the streamfunction is released from the ridge to the west when the negative (positive) anomaly of the upper-layer thickness reaches the western edge of the ridge from the east (see, e.g., a and b both in the right and left panels). This implies that the anomaly of the upperlayer thickness above the ridge carries the information of the volume transport variation to the west of the ridge. In the next subsection, we show a physical interpretation for the generation and behavior of the anomaly of the upper-layer thickness, and show why such baroclinic activity is related to the anomaly of the volume transport. b. Physical interpretation In order to interpret the behavior of the upper-layer thickness, we further simplify the governing equations. As shown in Fig. 9, the anomaly of the upper-layer thickness is mostly generated apart from the sidewall (see a on day 360 or b on day 180), and presumably not generated by the process in the sidewall boundary layer. Except in the sidewall boundary layer, the Coriolis and pressure gradient terms are dominant in Eqs. (2) and (3) throughout the year (not shown). Thus, as a consequence of the geostrophy, Eq. (4) can be rewritten as follows (see appendix B for details): h1 h1 h1 h1 dh ũ u 2, (11) t x y H dx where characteristic velocities, ũ and, are defined as ũ R. (12) H y H x As shown in Eq. (11), two processes govern the temporal variation of the upper-layer thickness. One is expressed by the second and third terms on the left-hand side of Eq. (11), which are the advection of the upperlayer thickness due to the combination of the vertically averaged flow and westward propagation of the long baroclinic Rossby wave [see Eq. (12)]. The advection due to the vertically averaged flow occurs when nonlinearity remains in Eq. (4). The right-hand side of Eq. (11) expresses another process that generates the anom-

12 MARCH 2002 ISOBE AND IMAWAKI 1005 FIG. 11. Horizontal distributions of the characteristic velocity above the ridge displayed every 60 days. The speed is indicated by gradation and the direction is indicated by arrows. Contours near the sidewalls are omitted. aly of the upper-layer thickness when the lower-layer flow impinges on the bottom slope. When the lowerlayer flow is heading toward the shallow (deep) portion, upwelling (downwelling) occurs through this term. Hereafter, we call this the impinging process as described by Isobe (2000). The ridge extends meridionally in this model, so that the impinging process is induced only by the zonal flow in the lower layer. Next, we show how these two processes work around the ridge. Figure 10 shows the distribution of the impinging term above the ridge (1550 km x 2050 km) every 60 days. As mentioned previously, the lower-layer flow remains throughout the annual cycle because isostasy is not accomplished for such short period variation. This means that the stratification acts as a high-pass filter on motion in the lower layer. The ridge prohibits motion from propagating westward due to the topographic effect so that the lower-layer flow impinging on the bottom slope is restricted to the eastern side of the ridge. Therefore, the impinging process works mainly on the eastern side of the ridge as shown in Fig. 10. Hence, the anomaly of the upper-layer thickness is generated on the eastern side of the ridge. Figure 11 shows the distribution of the characteristic velocities above the ridge every 60 days, representing the advection process of the upper-layer thickness. In general, the anomaly of the upper-layer thickness generated on the eastern side of the ridge is carried westward above the ridge. In the northern part, it is found that magnitudes of the characteristic velocities are small, and directions change both in time and space. This is because directions of two components of the characteristic velocity (i.e., the vertically averaged flow and westward propagation of the long baroclinic Rossby wave) are opposite in the northern part of the ridge. Namely, the propagation speed of the anomaly of the upper-layer thickness varies meridionally. The anomaly generated in the southern part of the ridge quickly moves westward, while the anomaly in the northern part is likely to move slowly. Thereafter, the anomaly of the upper-layer thickness reaches the western half of the ridge. Figure 12 shows the speed of the lower-layer flow above the ridge every 60 days. The lower-layer flow almost vanishes above the western half of the ridge because the topographic effect prohibits lower-layer motion from crossing the ridge and because the steady-state solution with isostasy dominates the western half of the ridge. As shown in Fig. 12, large lower-layer flow is found along geostrophic contours of the lower layer east of the ridge (see lower-left panel of Fig. 4). Figure 13 shows the space time plot of the speed of the lower-layer flow along line A B parallel to the geostrophic contours. The speed of the lower-layer flow is large both in summer

13 1006 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32 FIG. 12. Horizontal distributions of the speed of the lower-layer flow above the ridge displayed every 60 days. Contour interval is 2 cm s 1. Also shown is the contour of 1 cm s 1 by the broken line. FIG. 13. Space time plot of the speed of the lower-layer flow along the line A B over the ridge, whose location is shown in the right panel with horizontal distribution of the speed on day 360. Contour interval is 2 cm s 1. Stippling has been chosen to emphasize speed greater than 6 cm s 1. and winter because the anomaly of the volume transport is large in these seasons. It is found that lower-layer motion generated around A propagates toward the southern sidewall with a timescale of about 50 days. In addition, lower-layer motion can be seen over the western half of the ridge although its speed is considerably weak. This motion also propagates toward the northern sidewall with a similar timescale (not shown). Thus, lower-layer motion generating the anomaly of the upperlayer thickness quickly propagates toward the sidewall, and decays afterwards. Consequently, the anomaly of the upper-layer thickness, generated by lower-layer motion east of the ridge, is not accompanied with lower-layer motion when it reaches the western edge of the ridge. Namely, isostasy is accomplished with respect to each anomaly of the upper-layer thickness moving over the bottom topography. As a result, the positive (negative) anomaly of the upper-layer thickness carries the information of the positive (negative) anomaly of the volume transport when it reaches the western edge of the ridge. This experiment shows that the annual range of the volume transport variation is about 10 Sv west of the ridge. The variation is caused mostly by the baroclinic activity generated on the ridge. As mentioned previously, the time when the anomaly reaches the western edge of the ridge varies meridionally because of the difference of the characteristic velocities. Hence, the

14 MARCH 2002 ISOBE AND IMAWAKI 1007 phase of the annual variation of the volume transport is delayed northward as shown in Fig. 8. c. Comparison with the ASUKA observations As mentioned previously, the Kuroshio transport in the present study is defined as the volume transport between the western boundary and western edge of the ridge, corresponding to that of the Kuroshio throughflow across the whole ASUKA line. We here compare the present model results with observed values in Fig. 1, although rigorous comparison should be done using a more realistic ocean general circulation model. As shown in Fig. 1, the observed Kuroshio transport across the ASUKA line has an annual range of about 10 Sv, which is one-fourth or one-fifth of the Sverdrup transport estimated from ECMWF wind data. In our model, the annual range of the wind stress curl integrated zonally is adjusted to be the same as that estimated by the observed wind. The resultant annual range of the volume transport west of the ridge is about 10 Sv in the model (see Fig. 6), which is nearly the same as the observed value despite the simplified topography of the model. Considering that a one-layer model with realistic topography is unlikely to reproduce the annual range of the Kuroshio volume transport (e.g., Kubota et al. 1995), it is suggested that the annual range of the Kuroshio volume transport is mostly determined by the baroclinic activity generated above the ridge. In the present model, the phase of the annual variation of the Kuroshio transport is delayed about one month in comparison with that of the Sverdrup transport. However, such a delay is not clear in the observed transport in Fig. 1. We wish to emphasize that the phase of the annual variation of the volume transport varies in latitude as shown in Fig. 8. This means that the phase of the annual variation of the observed Kuroshio transport cannot be reproduced by a model with a simplified topography and wind field. d. Kuroshio transport across the PN-line In the present study, our concern is mostly focused on the annual variation in the volume transport of the Kuroshio throughflow across the ASUKA line. However, previous observational and numerical studies often focus on the eastward Kuroshio transport southwest of Japan, especially across the PN line in the East China Sea (see Fig. 1 for its location). For instance, based on the geostrophic calculation referred to the observed surface current velocity, Ichikawa and Chaen (2000) show that the Kuroshio transport across the PN line is maximum in summer and minimum in winter with an annual range of 14 Sv. We here consider how the annual range of the Kuroshio transport is determined for the PN line. Besides the annual variation controlled around the Izu Ogasawara Ridge, annual variation of the Kuroshio transport across the PN line may contain the variation of the closed circulation in the Philippine Basin. It is reasonable to consider that the annual variation of the Sverdrup transport integrated over the Philippine Basin is compensated by the eastward Kuroshio transport southwest of Japan. However, according to the numerical modeling by Sekine and Kutsuwada (1994), such a variation is blocked by the shelf slope east of Ryukyu Islands, and cannot propagate into the East China Sea. Rather, local downstream wind stress is responsible for the annual variation of the Kuroshio transport across the PN line (Ichikawa and Beardsley 1993). Besides the local wind forcing, the baroclinic activity may enhance the annual range of the Kuroshio transport across the PN line. Using a realistic ocean general circulation model, Kagimoto and Yamagata (1997) reproduce the annual variation of the Kuroshio transport across the PN line. They show the large contribution of an anticyclonic eddy to the Kuroshio transport variation. Their study suggests that the annual variation of the Kuroshio transport is controlled by baroclinic activity around the Ryukyu Islands. e. Variation with other timescales It is of interest to discuss the response of the simplified ocean model with the ridge to winds with various periods although this was not the main subject of this study. Even if the wind varies on very short timescales (less than annual), the impinging process occurs above the ridge because the lower-layer flow is still found at higher frequencies. However, the anomaly of the upperlayer thickness does not become large because upwelling and downwelling stop at once as the wind varies rapidly. Thus, the baroclinic activity does not work effectively. As a result, the meridional ridge acts as a barrier to the westward propagation of the volume transport variation. If the timescale of the wind variation is much longer than the time taken for the long baroclinic Rossby wave to cross the ocean, the baroclinic activity again does not work effectively above the ridge because isostasy is accomplished even east of the ridge. The baroclinic activity on the ridge, modifying the amplitude and phase of the volume transport variation, occurs for timescales up to several years. In fact, decadal variation of the sea level in the Gulf Stream region, although not the volume transport, is well reproduced by simple numerical models in which baroclinic Rossby waves are forced by the observed wind (Sturges and Hong 1995; Sturges et al. 1998; Hong et al. 2000). 5. Conclusions In the present study, we identify the baroclinic activity inducing the annual variation of the Kuroshio volume transport observed on the ASUKA line. The scenario is summarized as follows. For annual variation, lower-layer motion is found in areas surrounding the ridge because isostasy is not

15 1008 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32 achieved within such a short timescale. Thus, the flow impinges on the bottom slope. Anomalies of the upperlayer thickness are generated on the eastern side of the ridge through the impinging process expressed by the term on the right-hand side of Eq. (11). Thereafter, anomalies move westward with characteristic velocities composed of the vertically averaged flow and westward propagation of the long baroclinic Rossby wave. As anomalies of the upper-layer thickness move westward above the ridge, isostasy is accomplished with respect to these anomalies. As a result, the positive (negative) anomaly of the upper-layer thickness carries information about the positive (negative) anomaly of the volume transport as it reaches the western edge of the ridge. Thereafter, anomalies of the volume transport are released to the west of the ridge. When modeling the annual variation of the Kuroshio volume transport, it is important to reproduce the baroclinic activity as well as the circulation induced by the local wind forcing. Particularly, the accurate expression of the impinging process. [see Eq. (11)] is crucial because this process is a source of the baroclinic activity controlling the annual cycle of the Kuroshio volume transport. Acknowledgments. Authors express their sincere thanks to Drs. Yukio Masumoto and Hiroshi Uchida for their fruitful discussions. Thanks are also extended to Mr. Rolando S. Balotro for his critical reading of the manuscript. Detailed comments made by the editors and two anonymous reviewers helped to clarify many parts of an early version of the manuscript. This study has been supported by the Core Research for Evolutional Science and Technology (CREST) of Japan Science and Technology Corporation (JST). APPENDIX A Applicability of the Model Topography to the Real Ocean In the present study, we emphasize the impinging process as a source of the baroclinic activity above the ridge. As shown on the right-hand side of Eq. (11), the magnitude F of this process is measured by h1 H F u 2, H x where h 1, H, and u 2 are the upper-layer thickness, depth, and lower-layer velocity, respectively. Applying idealized topography in the present study [h m, H 3000 m, H/ x 1500 m/(500 km/2)], F above the ridge is estimated to be u 2. As shown in Fig. 1, the real Izu Ogasawara Ridge is narrower and steeper than the modeled ridge (W 200 km; the ridge with height 2000 m in a flat-bottom region with depth 5000 m). A narrow ridge requires fine resolution of the model, so we adopted a relatively gentle slope. In addition, the lower-layer flow would be weakened in the flat-bottom region with increased depth of 5000 m. As shown in the text, the lower-layer flow is the time-dependent part of the barotropic flow (i.e., vertically homogeneous flow). Even though the deeper flatbottom region is adopted in the model, time-dependent part of the volume transport does not change because it is determined by the wind stress. Therefore, multiplying the ratio of the model to real depths, the real lower-layer flow u 2 is evaluated to be 0.6u 2. Thus, a plausible value for F for realistic topography [h m, H 5000 m, H/ x 2000 m/(200 km/2), and u 2 ] is estimated to be u 2. This value is nearly the same as that for the modeled topography. The values of F are nearly the same even if H is replaced with the mean depth on the slope. Therefore, the modeled modification of the topography likely has little effect on the impinging process. APPENDIX B Simplification of the Governing Equations Assuming geostrophy, Eqs. (2) and (3) can be written as h f ( ) g x h f (u u ) g (B1) y Using Eqs. (6) and (B1), the velocities in the upper layer are obtained as 1 g h2 h1 u1 H y fh y 1 g h2 h1 1. (B2) H x fh x Substituting the above velocities into Eq. (4) yields h1 1 h1 1 h 2 1 R t H y x H x y h1 1 g h1 h1 dh, (B3) H H y fh y dx where R 2 is the internal Rossby deformation radius squared ( g h 1 h 2 / f 2 H). Using Eqs. (6) and (B1), the lower-layer velocity in the x direction (u 2 ) is expressed as 1 g h1 h1 u2. (B4) H y fh y Substituting Eq. (B4) into the parenthesis on the righthand side of Eq. (B3), we have Eq. (11). REFERENCES Anderson, D. L. T., and P. Killworth, 1977: Spin-up of a stratified ocean, with topography. Deep-Sea Res., 24,

16 MARCH 2002 ISOBE AND IMAWAKI 1009, and R. A. Corry, 1985: Seasonal transport variations in the Florida Straits: A model study. J. Phys. Oceanogr., 15, Barnier, B., 1988: Numerical study on the influence of the Mid- Atlantic Ridge on nonlinear first-mode baroclinic Rossby waves generated by seasonal winds. J. Phys. Oceanogr., 18, Gerdes, R., and C. Wübber, 1991: Seasonal variability of the North Atlantic Ocean A model intercomparison. J. Phys. Oceanogr., 21, Gill, A. E., and P. P. Niiler, 1973: The theory of the seasonal variability in the ocean. Deep-Sea Res., 20, Greatbatch, R. J., and A. Goulding, 1989: Seasonal variations in a linear barotropic model of the North Atlantic driven by the Hellerman and Rosenstein wind stress field. J. Phys. Oceanogr., 19, , and, 1990: On the seasonal variation of transport through the Tokara Strait. J. Oceanogr. Soc. Japan, 46, Hasunuma, K., and K. Yoshida, 1978: Splitting of the subtropical gyre in the western North Pacific. J. Oceanogr. Soc. Japan, 37, Hong, B. G, W. Sturges, and A. J. Clarke, 2000: Sea level on the U.S. East Coast: Decadal variability caused by open ocean windcurl forcing. J. Phys. Oceanogr., 30, Ichikawa, H., and R. C. Beardsley, 1993: Temporal and spatial variability of volume transport of the Kuroshio in the East China Sea. Deep-Sea Res., 40, , and M. Chaen, 2000: Seasonal variation of heat and freshwater transports by the Kuroshio in the East China Sea. J. Mar. Sys., 24, Imawaki, S., H. Uchida, H. Ichikawa, M. Fukasawa, S. Umatani, and ASUKA Group,1997: Time series of the Kuroshio transport derived from field observation and altimetry data. International WOCE Newsletter, Vol. 25, ,,,,, and, 2001: Satellite altimeter monitoring the Kuroshio transport south of Japan. Geophys. Res. Lett., 28, Isobe, A., 2000: Two-layer model on the branching of the Kuroshio southwest of Kyushu, Japan. J. Phys. Oceanogr., 30, Jarvis, R. A., and G. Veronis, 1994: Strong deep recirculations in a two-layer wind driven ocean. J. Phys. Oceanogr., 24, Kagimoto, T., and T. Yamagata, 1997: Seasonal transport variations of the Kuroshio: An OGCM simulation. J. Phys. Oceanogr., 27, Kubota, M., H. Yokota, and T. Okamoto, 1995: Mechanism of the seasonal transport variation through the Tokara Strait. J. Oceanogr., 51, Large, W. G., and S. Pond, 1981: Open ocean momentum flux measurements in moderate to strong winds, J. Phys. Oceanogr., 11, Molinari, R. L., W. D. Wilson, and K. D. Leaman, 1985: Volume and heat transports of the Florida Current: April 1982 to August Science, 227, Niiler, P. P., and W. S. Richardson, 1973: Seasonal variability of the Florida Current. J. Mar. Res., 31, Rhines, P. B., and W. R. Young, 1982: A theory of wind-driven circulation. I.Mid-ocean gyres. J. Mar. Res., 40 (Suppl.), Rooth, C., H. Stommel, and G. Veronis, 1978: On motions in steady, layered, geostrophic models. J. Oceanogr. Soc. Japan, 34, Sato, O. T., and T. Rossby, 1995: Seasonal and low frequency variations in dynamic height anomaly and transport of the Gulf Stream. Deep-Sea Res., 42, Schott, F. A., and R. J. Zantopp, 1985: On the seasonal and interannual variability of the Florida Current: Seasonal and interannual variability. Science, 227, Sekine, Y., and K. Kutsuwada, 1994: Seasonal variation in volume transport of the Kuroshio south of Japan. J. Phys. Oceanogr., 24, Sturges, W., and B. G. Hong, 1995: Wind forcing of the Atlantic thermocline along 32 N at low frequencies. J. Phys. Oceanogr., 25, ,, and A. J. Clarke, 1998: Decadal wind forcing of the North Atlantic subtropical gyre. J. Phys. Oceanogr., 28, Thompson, K. R., J. R. N. Lazier, and B. Taylor, 1986: Wind-forced changes in Labrador Current transport. J. Geophys. Res., 91,

Stationary Rossby Waves and Shocks on the Sverdrup Coordinate

Stationary Rossby Waves and Shocks on the Sverdrup Coordinate Journal of Oceanography Vol. 51, pp. 207 to 224. 1995 Stationary Rossby Waves and Shocks on the Sverdrup Coordinate ATSUSHI KUBOKAWA Graduate School of Environmental Earth Science, Hokkaido University,

More information

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

The Taiwan-Tsushima Warm Current System: Its Path and the Transformation of the Water Mass in the East China Sea Journal of Oceanography, Vol. 55, pp. 185 to 195. 1999 The Taiwan-Tsushima Warm Current System: Its Path and the Transformation of the Water Mass in the East China Sea ATSUHIKO ISOBE Department of Earth

More information

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

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

More information

Depth Distribution of the Subtropical Gyre in the North Pacific

Depth Distribution of the Subtropical Gyre in the North Pacific Journal of Oceanography, Vol. 58, pp. 525 to 529, 2002 Short Contribution Depth Distribution of the Subtropical Gyre in the North Pacific TANGDONG QU* International Pacific Research Center, SOEST, University

More information

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

The Effect of Koshu Seamount on the Formation of the Kuroshio Large Meander South of Japan 1624 J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y VOLUME 41 The Effect of Koshu Seamount on the Formation of the Kuroshio Large Meander South of Japan TAKAHIRO ENDOH Research Institute for

More information

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

Driving Mechanism of Band Structure of Mean Current over the Continental Shelf 1839 Driving Mechanism of Band Structure of Mean Current over the Continental Shelf ATSUHIKO ISOBE Department of Earth System Science and Technology, Interdisciplinary Graduate School of Engineering Sciences,

More information

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

General Comment on Lab Reports: v. good + corresponds to a lab report that: has structure (Intro., Method, Results, Discussion, an Abstract would be General Comment on Lab Reports: v. good + corresponds to a lab report that: has structure (Intro., Method, Results, Discussion, an Abstract would be a bonus) is well written (take your time to edit) shows

More information

Lecture 8. Lecture 1. Wind-driven gyres. Ekman transport and Ekman pumping in a typical ocean basin. VEk

Lecture 8. Lecture 1. Wind-driven gyres. Ekman transport and Ekman pumping in a typical ocean basin. VEk Lecture 8 Lecture 1 Wind-driven gyres Ekman transport and Ekman pumping in a typical ocean basin. VEk wek > 0 VEk wek < 0 VEk 1 8.1 Vorticity and circulation The vorticity of a parcel is a measure of its

More information

Islands in Zonal Flow*

Islands in Zonal Flow* 689 Islands in Zonal Flow* MICHAEL A. SPALL Department of Physical Oceanography, Woods Hole Oceanographic Institution, Woods Hole, Massachusetts (Manuscript received 1 April 003, in final form 9 June 003)

More information

Internal boundary layers in the ocean circulation

Internal boundary layers in the ocean circulation Internal boundary layers in the ocean circulation Lecture 9 by Andrew Wells We have so far considered boundary layers adjacent to physical boundaries. However, it is also possible to find boundary layers

More information

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.

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. Wind Gyres Here we derive the simplest (and oldest; Stommel, 1948) theory to explain western boundary currents like the Gulf Stream, and then discuss the relation of the theory to more realistic gyres.

More information

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

Note that Rossby waves are tranverse waves, that is the particles move perpendicular to the direction of propagation. f up, down (clockwise) Ocean 423 Rossby waves 1 Rossby waves: Restoring force is the north-south gradient of background potential vorticity (f/h). That gradient can be due to either the variation in f with latitude, or to a

More information

SIO 210: Dynamics VI (Potential vorticity) L. Talley Fall, 2014 (Section 2: including some derivations) (this lecture was not given in 2015)

SIO 210: Dynamics VI (Potential vorticity) L. Talley Fall, 2014 (Section 2: including some derivations) (this lecture was not given in 2015) SIO 210: Dynamics VI (Potential vorticity) L. Talley Fall, 2014 (Section 2: including some derivations) (this lecture was not given in 2015) Variation of Coriolis with latitude: β Vorticity Potential vorticity

More information

Upper Ocean Circulation

Upper Ocean Circulation Upper Ocean Circulation C. Chen General Physical Oceanography MAR 555 School for Marine Sciences and Technology Umass-Dartmouth 1 MAR555 Lecture 4: The Upper Oceanic Circulation The Oceanic Circulation

More information

Meridional circulation in the western coastal zone: Qinyan Liu +$ & Rui Xin Huang +* Guangzhou, China. February 2, 2010

Meridional circulation in the western coastal zone: Qinyan Liu +$ & Rui Xin Huang +* Guangzhou, China. February 2, 2010 Meridional circulation in the western coastal zone: II. The regulation by pressure gradient set up through basin scale circulation and the western boundary current transport Qinyan Liu +$ & Rui Xin Huang

More information

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

Sensitivity of the Interannual Kuroshio Transport Variation South of Japan to Wind Dataset in OGCM Calculation Journal of Oceanography, Vol. 60, pp. 341 to 350, 2004 Sensitivity of the Interannual Kuroshio Transport Variation South of Japan to Wind Dataset in OGCM Calculation HIROSHI YOSHINARI 1 *, MOTOYOSHI IKEDA

More information

Equilibrium surface elevation distribution in a baroclinic ocean with realistic bottom topography

Equilibrium surface elevation distribution in a baroclinic ocean with realistic bottom topography JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 08, NO. C0, 330, doi:0.029/2002jc00579, 2003 Equilibrium surface elevation distribution in a baroclinic ocean with realistic bottom topography Srdjan Dobricic Institute

More information

1. The figure shows sea surface height (SSH) anomaly at 24 S (southern hemisphere), from a satellite altimeter.

1. The figure shows sea surface height (SSH) anomaly at 24 S (southern hemisphere), from a satellite altimeter. SIO 210 Problem Set 3 November 16, 2015 1. The figure shows sea surface height (SSH) anomaly at 24 S (southern hemisphere), from a satellite altimeter. (a) What is the name of this type of data display?_hovmöller

More information

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

Mean Stream-Coordinate Structure of the Kuroshio Extension First Meander Trough Mean Stream-Coordinate Structure of the Kuroshio Extension First Meander Trough 6 March, 2008 Penelope J. Howe, Kathleen A. Donohue, and D. Randolph Watts Graduate School of Oceanography University of

More information

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

Baroclinic Rossby waves in the ocean: normal modes, phase speeds and instability Baroclinic Rossby waves in the ocean: normal modes, phase speeds and instability J. H. LaCasce, University of Oslo J. Pedlosky, Woods Hole Oceanographic Institution P. E. Isachsen, Norwegian Meteorological

More information

isopycnal outcrop w < 0 (downwelling), v < 0 L.I. V. P.

isopycnal outcrop w < 0 (downwelling), v < 0 L.I. V. P. Ocean 423 Vertical circulation 1 When we are thinking about how the density, temperature and salinity structure is set in the ocean, there are different processes at work depending on where in the water

More information

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

Lecture 14. Equations of Motion Currents With Friction Sverdrup, Stommel, and Munk Solutions Remember that Ekman's solution for wind-induced transport Lecture 14. Equations of Motion Currents With Friction Sverdrup, Stommel, and Munk Solutions Remember that Ekman's solution for wind-induced transport is which can also be written as (14.1) i.e., #Q x,y

More information

Atmosphere, Ocean and Climate Dynamics Fall 2008

Atmosphere, Ocean and Climate Dynamics Fall 2008 MIT OpenCourseWare http://ocw.mit.edu 12.003 Atmosphere, Ocean and Climate Dynamics Fall 2008 For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms. Problem

More information

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

Eddy Formation Near the Izu-Ogasawara Ridge and its Link with Seasonal Adjustment of the Subtropical Gyre in the Pacific 000 Journal of the Korean Society of Oceanography Vol. 37, No. 3, pp. 000 September 2002 INFORMATION Eddy Formation Near the Izu-Ogasawara Ridge and its Link with Seasonal Adjustment of the Subtropical

More information

Circulation in the South China Sea in summer of 1998

Circulation in the South China Sea in summer of 1998 Circulation in the South China Sea in summer of 1998 LIU Yonggang, YUAN Yaochu, SU Jilan & JIANG Jingzhong Second Institute of Oceanography, State Oceanic Administration (SOA), Hangzhou 310012, China;

More information

Measuring the Flow Through the Kerama Gap

Measuring the Flow Through the Kerama Gap DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. Measuring the Flow Through the Kerama Gap Mark Wimbush & Jae-Hun Park Graduate School of Oceanography University of Rhode

More information

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

A NUMERICAL EXPERIMENT OF 50-DAY RESONANCE INDUCED BY INDIAN OCEAN KELVIN WAVE IN THE SULAWESI SEA A NUMERICAL EXPERIMENT OF 50-DAY RESONANCE INDUCED BY INDIAN OCEAN KELVIN WAVE IN THE SULAWESI SEA Fadli Syamsudin Center for Research and Development of Technology for Natural Resources Inventory, Agency

More information

The Kuroshio East of Taiwan and in the East China Sea and the Currents East of Ryukyu Islands during Early Summer of 1996

The Kuroshio East of Taiwan and in the East China Sea and the Currents East of Ryukyu Islands during Early Summer of 1996 Journal of Oceanography, Vol. 54, pp. 217 to 226. 1998 The Kuroshio East of Taiwan and in the East China Sea and the Currents East of Ryukyu Islands during Early Summer of 1996 YAOCHU YUAN 1, ARATA KANEKO

More information

Climate Variability Inferred from a Layered Model of the Ventilated Thermocline*

Climate Variability Inferred from a Layered Model of the Ventilated Thermocline* APRIL 1999 HUANG AND PEDLOSKY 779 Climate Variability Inferred from a Layered Model of the Ventilated Thermocline* RUI XIN HUANG AND JOSEPH PEDLOSKY Department of Physical Oceanography, Woods Hole Oceanographic

More information

Lecture 2 ENSO toy models

Lecture 2 ENSO toy models Lecture 2 ENSO toy models Eli Tziperman 2.3 A heuristic derivation of a delayed oscillator equation Let us consider first a heuristic derivation of an equation for the sea surface temperature in the East

More information

On the formation of Subtropical Countercurrent to the west of the Hawaiian Islands

On the formation of Subtropical Countercurrent to the west of the Hawaiian Islands JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. C5, 3167, doi:10.1029/2002jc001366, 2003 On the formation of Subtropical Countercurrent to the west of the Hawaiian Islands Qinyu Liu, Shaoxia Wang, Qi Wang,

More information

I. Variability of off-shore thermocline depth and western boundary current. Rui Xin Huang *+ & Qinyan Liu +$

I. Variability of off-shore thermocline depth and western boundary current. Rui Xin Huang *+ & Qinyan Liu +$ Meridional circulation in the western coastal zone: I. Variability of off-shore thermocline depth and western boundary current transport on interannual and decadal time scales Rui Xin Huang *+ & Qinyan

More information

Characteristics of Sea Surface Circulation and Eddy Field in the South China Sea Revealed by Satellite Altimetric Data

Characteristics of Sea Surface Circulation and Eddy Field in the South China Sea Revealed by Satellite Altimetric Data Journal of Oceanography, Vol. 56, pp. 331 to 344, 2000 Characteristics of Sea Surface Circulation and Eddy Field in the South China Sea Revealed by Satellite Altimetric Data AKIHIKO MORIMOTO 1 *, KOICHI

More information

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

P-Vector Inverse Method Evaluated Using the Modular Ocean Model (MOM) Journal of Oceanography, Vol. 54, pp. 185 to 198. 1998 P-Vector Inverse Method Evaluated Using the Modular Ocean Model (MOM) PETER C. CHU 1, CHENWU FAN 1 and WENJU CAI 2 1 Naval Postgraduate School, Monterey,

More information

Buoyancy-forced circulations in shallow marginal seas

Buoyancy-forced circulations in shallow marginal seas Journal of Marine Research, 63, 729 752, 2005 Buoyancy-forced circulations in shallow marginal seas by Michael A. Spall 1 ABSTRACT The properties of water mass transformation and the thermohaline circulation

More information

Water mass transport associated with the oceanic fronts in the northwestern Pacific Ocean HIDEYUKI NAKANO (METEOROLOGICAL RESEARCH INSTITUTE)

Water mass transport associated with the oceanic fronts in the northwestern Pacific Ocean HIDEYUKI NAKANO (METEOROLOGICAL RESEARCH INSTITUTE) Water mass transport associated with the oceanic fronts in the northwestern Pacific Ocean HIDEYUKI NAKANO (METEOROLOGICAL RESEARCH INSTITUTE) How is the Kuroshio-origin water distributed in the subtropical

More information

Large-Scale Circulations Forced by Localized Mixing over a Sloping Bottom*

Large-Scale Circulations Forced by Localized Mixing over a Sloping Bottom* AUGUST 001 SPALL 369 Large-Scale Circulations Forced by Localized Miing over a Sloping Bottom* MICHAEL A. SPALL Department of Physical Oceanography, Woods Hole Oceanographic Institution, Woods Hole, Massachusetts

More information

Why Western Boundary Currents in Realistic Oceans are Inviscid: A Link between Form Stress and Bottom Pressure Torques

Why Western Boundary Currents in Realistic Oceans are Inviscid: A Link between Form Stress and Bottom Pressure Torques 2871 Why Western Boundary Currents in Realistic Oceans are Inviscid: A Link between Form Stress and Bottom Pressure Torques CHRIS W. HUGHES Proudman Oceanographic Laboratory, Bidston Observatory, Prenton,

More information

The Planetary Circulation System

The Planetary Circulation System 12 The Planetary Circulation System Learning Goals After studying this chapter, students should be able to: 1. describe and account for the global patterns of pressure, wind patterns and ocean currents

More information

Exam Questions & Problems

Exam Questions & Problems 1 Exam Questions & Problems Summer School on Dynamics of the North Indian Ocean National Institute of Oceanography, Dona Paula, Goa General topics that have been considered during this course are indicated

More information

An Introduction to Coupled Models of the Atmosphere Ocean System

An Introduction to Coupled Models of the Atmosphere Ocean System An Introduction to Coupled Models of the Atmosphere Ocean System Jonathon S. Wright jswright@tsinghua.edu.cn Atmosphere Ocean Coupling 1. Important to climate on a wide range of time scales Diurnal to

More information

Ocean Dynamics. The Great Wave off Kanagawa Hokusai

Ocean Dynamics. The Great Wave off Kanagawa Hokusai Ocean Dynamics The Great Wave off Kanagawa Hokusai LO: integrate relevant oceanographic processes with factors influencing survival and growth of fish larvae Physics Determining Ocean Dynamics 1. Conservation

More information

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

Lecture #2 Planetary Wave Models. Charles McLandress (Banff Summer School 7-13 May 2005) Lecture #2 Planetary Wave Models Charles McLandress (Banff Summer School 7-13 May 2005) 1 Outline of Lecture 1. Observational motivation 2. Forced planetary waves in the stratosphere 3. Traveling planetary

More information

Ocean surface circulation

Ocean surface circulation Ocean surface circulation Recall from Last Time The three drivers of atmospheric circulation we discussed: Differential heating Pressure gradients Earth s rotation (Coriolis) Last two show up as direct

More information

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

Eddy-resolving Simulation of the World Ocean Circulation by using MOM3-based OGCM Code (OFES) Optimized for the Earth Simulator Chapter 1 Atmospheric and Oceanic Simulation Eddy-resolving Simulation of the World Ocean Circulation by using MOM3-based OGCM Code (OFES) Optimized for the Earth Simulator Group Representative Hideharu

More information

ROSSBY WAVE PROPAGATION

ROSSBY WAVE PROPAGATION ROSSBY WAVE PROPAGATION (PHH lecture 4) The presence of a gradient of PV (or q.-g. p.v.) allows slow wave motions generally called Rossby waves These waves arise through the Rossby restoration mechanism,

More information

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

Characteristics of Variations of Water Properties and Density Structure around the Kuroshio in the East China Sea Journal of Oceanography, Vol. 54, pp. 605 to 617. 1998 Characteristics of Variations of Water Properties and Density Structure around the Kuroshio in the East China Sea EITAROU OKA and MASAKI KAWABE Ocean

More information

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

2/15/2012. Earth System Science II EES 717 Spring 2012 Earth System Science II EES 717 Spring 2012 1. The Earth Interior Mantle Convection & Plate Tectonics 2. The Atmosphere - Climate Models, Climate Change and Feedback Processes 3. The Oceans Circulation;

More information

6 Two-layer shallow water theory.

6 Two-layer shallow water theory. 6 Two-layer shallow water theory. Wewillnowgoontolookatashallowwatersystemthathastwolayersofdifferent density. This is the next level of complexity and a simple starting point for understanding the behaviour

More information

Measuring the Flow through the Kerama Gap

Measuring the Flow through the Kerama Gap DISTRIBUTION STATEMENT A. Approved for public release; distribution is unlimited. Measuring the Flow through the Kerama Gap Mark Wimbush & Jae-Hun Park Graduate School of Oceanography University of Rhode

More information

Eddy-induced meridional heat transport in the ocean

Eddy-induced meridional heat transport in the ocean GEOPHYSICAL RESEARCH LETTERS, VOL. 35, L20601, doi:10.1029/2008gl035490, 2008 Eddy-induced meridional heat transport in the ocean Denis L. Volkov, 1 Tong Lee, 1 and Lee-Lueng Fu 1 Received 28 July 2008;

More information

Boundary Layers: Homogeneous Ocean Circulation

Boundary Layers: Homogeneous Ocean Circulation Boundary Layers: Homogeneous Ocean irculation Lecture 7 by Angel Ruiz-Angulo The first explanation for the western intensification of the wind-driven ocean circulation was provided by Henry Stommel (948).

More information

Topographic Control of Basin and Channel Flows: The Role of Bottom Pressure Torques and Friction

Topographic Control of Basin and Channel Flows: The Role of Bottom Pressure Torques and Friction 1786 J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y VOLUME 36 Topographic Control of Basin and Channel Flows: The Role of Bottom Pressure Torques and Friction LAURA JACKSON* Department of Earth

More information

PAPER 333 FLUID DYNAMICS OF CLIMATE

PAPER 333 FLUID DYNAMICS OF CLIMATE MATHEMATICAL TRIPOS Part III Wednesday, 1 June, 2016 1:30 pm to 4:30 pm Draft 21 June, 2016 PAPER 333 FLUID DYNAMICS OF CLIMATE Attempt no more than THREE questions. There are FOUR questions in total.

More information

Wind- and buoyancy-forced upper ocean circulation in two-strait marginal seas with application to the Japan/East Sea

Wind- and buoyancy-forced upper ocean circulation in two-strait marginal seas with application to the Japan/East Sea JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. C1, 3006, 10.1029/2001JC000966, 2002 Wind- and buoyancy-forced upper ocean circulation in two-strait marginal seas with application to the Japan/East Sea

More information

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

Short-Range Prediction Experiments with Operational Data Assimilation System for the Kuroshio South of Japan Journal of Oceanography, Vol. 60, pp. 269 to 282, 2004 Short-Range Prediction Experiments with Operational Data Assimilation System for the Kuroshio South of Japan MASAFUMI KAMACHI 1 *, TSURANE KURAGANO

More information

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

BALANCED FLOW: EXAMPLES (PHH lecture 3) Potential Vorticity in the real atmosphere. Potential temperature θ. Rossby Ertel potential vorticity BALANCED FLOW: EXAMPLES (PHH lecture 3) Potential Vorticity in the real atmosphere Need to introduce a new measure of the buoyancy Potential temperature θ In a compressible fluid, the relevant measure

More information

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

Wind-driven Western Boundary Ocean Currents in Terran and Superterran Exoplanets Wind-driven Western Boundary Ocean Currents in Terran and Superterran Exoplanets By Edwin Alfonso-Sosa, Ph.D. Ocean Physics Education, Inc. 10-Jul-2014 Introduction Simple models of oceanic general circulation

More information

Wind-Driven Transport Fluctuations through Drake Passage: A Southern Mode

Wind-Driven Transport Fluctuations through Drake Passage: A Southern Mode AUGUST 1999 HUGHES ET AL. 1971 Wind-Driven Transport Fluctuations through Drake Passage: A Southern Mode CHRIS W. HUGHES Proudman Oceanographic Laboratory, Bidston Observatory, Birkenhead, Merseyside,

More information

Ocean dynamics: the wind-driven circulation

Ocean dynamics: the wind-driven circulation Ocean dynamics: the wind-driven circulation Weston Anderson March 13, 2017 Contents 1 Introduction 1 2 The wind driven circulation (Ekman Transport) 3 3 Sverdrup flow 5 4 Western boundary currents (western

More information

CANER SERTKAN ( ) WIND DRIVEN OCEAN CIRCULATION

CANER SERTKAN ( ) WIND DRIVEN OCEAN CIRCULATION CANER SERTKAN (517101005) WIND DRIVEN OCEAN CIRCULATION INTRODUCTION The waters of the ocean are continually moving in powerful currents like the Gulf Stream, in large gyres, in feautures visible from

More information

OCN/ATM/ESS 587. The wind-driven ocean circulation. Friction and stress. The Ekman layer, top and bottom. Ekman pumping, Ekman suction

OCN/ATM/ESS 587. The wind-driven ocean circulation. Friction and stress. The Ekman layer, top and bottom. Ekman pumping, Ekman suction OCN/ATM/ESS 587 The wind-driven ocean circulation. Friction and stress The Ekman layer, top and bottom Ekman pumping, Ekman suction Westward intensification The wind-driven ocean. The major ocean gyres

More information

Island Wakes in Shallow Water

Island Wakes in Shallow Water Island Wakes in Shallow Water Changming Dong, James C. McWilliams, et al Institute of Geophysics and Planetary Physics, University of California, Los Angeles 1 ABSTRACT As a follow-up work of Dong et al

More information

Processes Coupling the Upper and Deep Ocean on the Continental Slope

Processes Coupling the Upper and Deep Ocean on the Continental Slope Processes Coupling the Upper and Deep Ocean on the Continental Slope D. Randolph Watts Graduate School of Oceanography University of Rhode Island South Ferry Road Narragansett, RI 02882 phone:(401) 874-6507;

More information

SIO 210 Problem Set 2 October 17, 2011 Due Oct. 24, 2011

SIO 210 Problem Set 2 October 17, 2011 Due Oct. 24, 2011 SIO 210 Problem Set 2 October 17, 2011 Due Oct. 24, 2011 1. The Pacific Ocean is approximately 10,000 km wide. Its upper layer (wind-driven gyre*) is approximately 1,000 m deep. Consider a west-to-east

More information

MERIDIONAL OVERTURNING CIRCULATION: SOME BASICS AND ITS MULTI-DECADAL VARIABILITY

MERIDIONAL OVERTURNING CIRCULATION: SOME BASICS AND ITS MULTI-DECADAL VARIABILITY MERIDIONAL OVERTURNING CIRCULATION: SOME BASICS AND ITS MULTI-DECADAL VARIABILITY Gokhan Danabasoglu National Center for Atmospheric Research OUTLINE: - Describe thermohaline and meridional overturning

More information

Ocean Boundary Currents Guiding Question: How do western boundary currents influence climate and ocean productivity?

Ocean Boundary Currents Guiding Question: How do western boundary currents influence climate and ocean productivity? Name: Date: TEACHER VERSION: Suggested Student Responses Included Ocean Boundary Currents Guiding Question: How do western boundary currents influence climate and ocean productivity? Introduction The circulation

More information

SIO 210: Dynamics VI: Potential vorticity

SIO 210: Dynamics VI: Potential vorticity SIO 210: Dynamics VI: Potential vorticity Variation of Coriolis with latitude: β Vorticity Potential vorticity Rossby waves READING: Review Section 7.2.3 Section 7.7.1 through 7.7.4 or Supplement S7.7

More information

Semiannual variation in the western tropical Pacific Ocean

Semiannual variation in the western tropical Pacific Ocean Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 35, L16602, doi:10.1029/2008gl035058, 2008 Semiannual variation in the western tropical Pacific Ocean Tangdong Qu, 1 Jianping Gan, 2 Akio

More information

Transport driven by eddy momentum fluxes in the Gulf Stream Extension region

Transport driven by eddy momentum fluxes in the Gulf Stream Extension region GEOPHYSICAL RESEARCH LETTERS, VOL. 37,, doi:10.1029/2010gl045473, 2010 Transport driven by eddy momentum fluxes in the Gulf Stream Extension region R. J. Greatbatch, 1 X. Zhai, 2 M. Claus, 1 L. Czeschel,

More information

Observation of Oceanic Structure around Tosa-Bae Southeast of Shikoku

Observation of Oceanic Structure around Tosa-Bae Southeast of Shikoku Journal of Oceanography Vol. 50, pp. 543 to 558. 1994 Observation of Oceanic Structure around Tosa-Bae Southeast of Shikoku YOSHIHIKO SEKINE, HARUKI OHWAKI and MOTOYA NAKAGAWA Institute of Oceanography,

More information

Wind-Driven Fluctuating Western Boundary Currents

Wind-Driven Fluctuating Western Boundary Currents VOLUME 28 JOURNAL OF PYSICAL OCEANOGRAPY NOVEMBER 998 Wind-Driven Fluctuating Western Boundary Currents STEFANO PIERINI Dipartimento di Fisica, Università dell Aquila, L Aquila, Italy (Manuscript received

More information

North Atlantic circulation in three simulations of 1/12, 1/25, and 1/50

North Atlantic circulation in three simulations of 1/12, 1/25, and 1/50 North Atlantic circulation in three simulations of 1/12, 1/2, and 1/ Xiaobiao Xu and Eric Chassignet Center for ocean-atmospheric prediction studies Florida State University Motivation Numerical models

More information

Fig. 1. Structure of the equatorial thermocline and circulation (Wyrtki and Kilonsky, 1984).

Fig. 1. Structure of the equatorial thermocline and circulation (Wyrtki and Kilonsky, 1984). Lecture 18. Equatorial thermocline 4/5/006 11:40 AM 1. Introduction We have discussed dynamics of western boundary layers in the subtropical and subpolar basin. In particular, inertial boundary layer plays

More information

3. Midlatitude Storm Tracks and the North Atlantic Oscillation

3. Midlatitude Storm Tracks and the North Atlantic Oscillation 3. Midlatitude Storm Tracks and the North Atlantic Oscillation Copyright 2006 Emily Shuckburgh, University of Cambridge. Not to be quoted or reproduced without permission. EFS 3/1 Review of key results

More information

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

Convection Induced by Cooling at One Side Wall in Two-Dimensional Non-Rotating Fluid Applicability to the Deep Pacific Circulation Journal of Oceanography Vol. 52, pp. 617 to 632. 1996 Convection Induced by Cooling at One Side Wall in Two-Dimensional Non-Rotating Fluid Applicability to the Deep Pacific Circulation ICHIRO ISHIKAWA

More information

Wind Driven Ocean Circulation

Wind Driven Ocean Circulation Chapter 11 Wind Driven Ocean Circulation What drives the ocean currents? At first, we might answer, the winds drive the circulation. But if we think more carefully about the question, we might not be so

More information

Equatorial Superrotation on Tidally Locked Exoplanets

Equatorial Superrotation on Tidally Locked Exoplanets Equatorial Superrotation on Tidally Locked Exoplanets Adam P. Showman University of Arizona Lorenzo M. Polvani Columbia University Synopsis Most 3D atmospheric circulation models of tidally locked exoplanets

More information

Lecture 1. Amplitude of the seasonal cycle in temperature

Lecture 1. Amplitude of the seasonal cycle in temperature Lecture 6 Lecture 1 Ocean circulation Forcing and large-scale features Amplitude of the seasonal cycle in temperature 1 Atmosphere and ocean heat transport Trenberth and Caron (2001) False-colour satellite

More information

CHAPTER 4. THE HADLEY CIRCULATION 59 smaller than that in midlatitudes. This is illustrated in Fig. 4.2 which shows the departures from zonal symmetry

CHAPTER 4. THE HADLEY CIRCULATION 59 smaller than that in midlatitudes. This is illustrated in Fig. 4.2 which shows the departures from zonal symmetry Chapter 4 THE HADLEY CIRCULATION The early work on the mean meridional circulation of the tropics was motivated by observations of the trade winds. Halley (1686) and Hadley (1735) concluded that the trade

More information

Lecture 28: A laboratory model of wind-driven ocean circulation

Lecture 28: A laboratory model of wind-driven ocean circulation Lecture 28: A laboratory model of wind-driven ocean circulation November 16, 2003 1 GFD Lab XIII: Wind-driven ocean gyres It is relatively straightforward to demonstrate the essential mechanism behind

More information

Where Does Dense Water Sink? A Subpolar Gyre Example*

Where Does Dense Water Sink? A Subpolar Gyre Example* 810 JOURNAL OF PHYSICAL OCEANOGRAPHY Where Does Dense Water Sink? A Subpolar Gyre Example* MICHAEL A. SPALL AND ROBERT S. PICKART Department of Physical Oceanography, Woods Hole Oceanographic Institution,

More information

Non-linear patterns of eddy kinetic energy in the Japan/East Sea

Non-linear patterns of eddy kinetic energy in the Japan/East Sea Non-linear patterns of eddy kinetic energy in the Japan/East Sea O.O. Trusenkova, D.D. Kaplunenko, S.Yu. Ladychenko, V.B. Lobanov V.I.Il ichev Pacific Oceanological Institute, FEB RAS Vladivostok, Russia

More information

Upper Ocean - Topographic Coupling in an Ocean Model with High Vertical Resolution

Upper Ocean - Topographic Coupling in an Ocean Model with High Vertical Resolution Upper Ocean - Topographic Coupling in an Ocean Model with High Vertical Resolution by 1 Harley E. Hurlburt, 1 Patrick J. Hogan, 1 E. Joseph Metzger, 2 Charles E. Tilburg and 1 Jay F. Shriver 1 Naval Research

More information

A Study on Residual Flow in the Gulf of Tongking

A Study on Residual Flow in the Gulf of Tongking Journal of Oceanography, Vol. 56, pp. 59 to 68. 2000 A Study on Residual Flow in the Gulf of Tongking DINH-VAN MANH 1 and TETSUO YANAGI 2 1 Department of Civil and Environmental Engineering, Ehime University,

More information

Gulf Stream Transport Variability at Periods of Decades

Gulf Stream Transport Variability at Periods of Decades 1304 JOURNAL OF PHYSICAL OCEANOGRAPHY Gulf Stream Transport Variability at Periods of Decades W. STURGES AND B. G. HONG* Department of Oceanography, The Florida State University, Tallahassee, Florida (Manuscript

More information

Ocean Mixing and Climate Change

Ocean Mixing and Climate Change Ocean Mixing and Climate Change Factors inducing seawater mixing Different densities Wind stirring Internal waves breaking Tidal Bottom topography Biogenic Mixing (??) In general, any motion favoring turbulent

More information

Goals of this Chapter

Goals of this Chapter Waves in the Atmosphere and Oceans Restoring Force Conservation of potential temperature in the presence of positive static stability internal gravity waves Conservation of potential vorticity in the presence

More information

Lecture 25: Ocean circulation: inferences from geostrophic and thermal wind balance

Lecture 25: Ocean circulation: inferences from geostrophic and thermal wind balance Lecture 25: Ocean circulation: inferences from geostrophic and thermal wind balance November 5, 2003 Today we are going to study vertical sections through the ocean and discuss what we can learn about

More information

Taiwan Current (Kuroshio) and Impinging Eddies

Taiwan Current (Kuroshio) and Impinging Eddies Journal of Oceanography, Vol. 55, pp. 609 to 617. 1999 Taiwan Current (Kuroshio) and Impinging Eddies YIH YANG 1, CHO-TENG LIU 1, JIAN-HWA HU 2 and MOMOKI KOGA 3 1 Institute of Oceanography, National Taiwan

More information

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

Comparison Figures from the New 22-Year Daily Eddy Dataset (January April 2015) Comparison Figures from the New 22-Year Daily Eddy Dataset (January 1993 - April 2015) The figures on the following pages were constructed from the new version of the eddy dataset that is available online

More information

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

1/27/2010. With this method, all filed variables are separated into. from the basic state: Assumptions 1: : the basic state variables must Lecture 5: Waves in Atmosphere Perturbation Method With this method, all filed variables are separated into two parts: (a) a basic state part and (b) a deviation from the basic state: Perturbation Method

More information

Four-Gyre Circulation in a Barotropic Model with Double-Gyre Wind Forcing

Four-Gyre Circulation in a Barotropic Model with Double-Gyre Wind Forcing 1461 Four-Gyre Circulation in a Barotropic Model with Double-Gyre Wind Forcing RICHARD J. GREATBATCH Department of Oceanography, Dalhousie University, Halifax, Nova Scotia, Canada B. T. NADIGA Earth and

More information

Chapter 9. Geostrophy, Quasi-Geostrophy and the Potential Vorticity Equation

Chapter 9. Geostrophy, Quasi-Geostrophy and the Potential Vorticity Equation Chapter 9 Geostrophy, Quasi-Geostrophy and the Potential Vorticity Equation 9.1 Geostrophy and scaling. We examined in the last chapter some consequences of the dynamical balances for low frequency, nearly

More information

NOTES AND CORRESPONDENCE. Effect of Sea Surface Temperature Wind Stress Coupling on Baroclinic Instability in the Ocean

NOTES AND CORRESPONDENCE. Effect of Sea Surface Temperature Wind Stress Coupling on Baroclinic Instability in the Ocean 1092 J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y VOLUME 37 NOTES AND CORRESPONDENCE Effect of Sea Surface Temperature Wind Stress Coupling on Baroclinic Instability in the Ocean MICHAEL A.

More information

Dynamics and Kinematics

Dynamics and Kinematics Geophysics Fluid Dynamics () Syllabus Course Time Lectures: Tu, Th 09:30-10:50 Discussion: 3315 Croul Hall Text Book J. R. Holton, "An introduction to Dynamic Meteorology", Academic Press (Ch. 1, 2, 3,

More information

C

C C 0.8 0.4 0.2 0.0-0.2-0.6 Fig. 1. SST-wind relation in the North Pacific and Atlantic Oceans. Left panel: COADS SST (color shade), surface wind vectors, and SLP regressed upon the Pacific Decadal Oscillation

More information

Geophysics Fluid Dynamics (ESS228)

Geophysics Fluid Dynamics (ESS228) Geophysics Fluid Dynamics (ESS228) Course Time Lectures: Tu, Th 09:30-10:50 Discussion: 3315 Croul Hall Text Book J. R. Holton, "An introduction to Dynamic Meteorology", Academic Press (Ch. 1, 2, 3, 4,

More information

Wind-driven flow over topography

Wind-driven flow over topography Journal of Marine Research, 74, 229 248, 26 Wind-driven flow over topography by Michael A. Spall,2 ABSTRACT The space and time scales over which wind forcing can directly drive flows over regions of closed

More information

SIO 210 Final examination Wednesday, December 12, :30-2:30 Eckart 227 Name:

SIO 210 Final examination Wednesday, December 12, :30-2:30 Eckart 227 Name: SIO 210 Final examination Wednesday, December 12, 2018 11:30-2:30 Eckart 227 Name: Please put your initials or name on each page, especially if you pull pages apart. Turn off all phones, ipods, etc. and

More information