<<

JANUARY 2002 QIU 353

Large-Scale Variability in the Midlatitude Subtropical and Subpolar North Paci®c Ocean: Observations and Causes

BO QIU Department of Oceanography, University of Hawaii at Manoa, Honolulu, Hawaii

(Manuscript received 26 January 2001, in ®nal form 1 June 2001)

ABSTRACT Altimetric data from the 8-yr TOPEX/Poseidon (T/P) mission (Oct 1992±Jul 2000) are used to investigate large-scale circulation changes in the three current systems of the midlatitude North Paci®c Ocean: the North Paci®c Current (NPC), the Alaska gyre, and the western subarctic gyre (WSG). To facilitate the understanding of the observed changes, a two-layer ocean model was adopted that includes ®rst-mode baroclinic Rossby wave dynamics and barotropic Sverdrup dynamics. The NPC intensi®ed steadily over the T/P period from 1992 to 1998. Much of this intensi®cation is due to the persistent sea surface height (SSH) drop on the northern side of the NPC. A similar SSH trend is also found in the interior of the Alaska gyre. Both of these SSH changes are shown to be the result of surface wind stress curl forcing accumulated along the baroclinic Rossby wave characteristics initiated from the eastern boundary. In addition to the interior SSH signals, the intensity of the Alaska gyre is shown to depend also on the SSH anomalies along the Canada/Alaska coast, and these anomalies are shown to be jointly determined by the signals propagating from lower latitudes and those forced locally by the alongshore surface winds. The WSG changed interannually from a zonally elongated gyre in 1993±95 to a zonally more contracted gyre in 1997±99. This structural change is due to the interannual SSH anomalies within the WSG as a result of the baroclinic Rossby wave adjustment attenuated by eddy dissipation. Along the western boundary of the subpolar North Paci®c, variability of the East (EKC) and Oyashio is in balance with that of the interior Sverdrup ¯ow on the annual and year-to-year timescales. On the multiyear timescales, the EKC/Oyashio variability is shown to be determined by the baroclinic SSH signals.

1. Introduction (Chelton and Davis 1982). After veering southwestward at the apex of the , the Alaska Current Large-scale interior circulations in the North Paci®c is renamed the Alaskan Stream. Flowing southwestward Ocean north of 32ЊN consist of four major current sys- along the Aleutian Islands, the Alaskan Stream consti- tems: the Kuroshio Extension, the North Paci®c Current tutes the western of the Alaska gyre (NPC), the Alaska gyre, and the western subarctic gyre (e.g., Favorite et al. 1976; Reed 1984; Musgrave et al. (WSG) (see Fig. 1). The Kuroshio Extension is the off- shore extension of the western boundary current of the 1992). North Paci®c subtropical gyre (Kawai 1972). It is an As the Alaskan Stream ¯ows farther westward, part eastward inertial jet characterized by large-amplitude of it enters the Bering Sea through deep passages along meanders and vigorous pinched-off eddies (e.g., Mizuno the Aleutian Islands between 168Њ and 172ЊE. The re- and White 1983). The Kuroshio Extension loses these maining continues westward along the southern edge of inertial current characteristics around the date line, and the Aleutian Islands and constitutes the northern limb thereafter it broadens to become the eastward ¯owing of the western subarctic gyre. Upon reaching the Kam- North Paci®c Current (White 1982). After encountering chatka peninsula, the Alaskan Stream turns southwest- the coast, the NPC bifurcates. In addition ward to form the western boundary current of the WSG: to forming the coastal , the southern the East Kamchatka Current and its southern continu- branch of the NPC veers southward and joins the interior ation, the Oyashio. The WSG is closed to the south by Sverdrup ¯ow of the subtropical gyre. The northern the Subarctic Current after the Oyashio veers eastward branch of the NPC forms the origin of the Alaska Cur- around 42ЊN, east of Hokkaido (Favorite et al. 1976). rent and constitutes the eastern limb of the Alaska gyre In contrast to this picture of the large-scale mean circulation, our knowledge about the variability in the midlatitude North Paci®c is more fragmentary. Al- Corresponding author address: Dr. Bo Qiu, Department of Ocean- though regional long-term measurements exist (a good ography, University of Hawaii at Manoa, 1000 Pope Road, Honolulu, HI 96822. example of this is the P-line of repeat hydrographic E-mail: [email protected] surveys across the Alaska Current; see Tabata 1991),

᭧ 2002 American Meteorological Society

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC 354 JOURNAL OF VOLUME 32

FIG. 1. (a) Surface dynamic topography (cm) relative to 1000 dbar calculated from the Levitus and Boyer (1994) and Levitus et al. (1994) climatology. (b) Areas de®ning the four current systems: the Kuroshio Extension (KE), the North Paci®c Current (NPC), the Alaska Gyre (AG), and the Western Subarctic Gyre (WSG). investigations that examine the individual current sys- recent studies (e.g., Fu and Davidson 1995; Fukumori tems as entities and explore their interconnections have et al. 1998). In this study, we will instead focus on the been lacking. This is primarily due to the vast geo- SSH signals that have timescales longer than O(100 graphical extent of each current system. The launch of days). It is worth emphasizing that the longer time series the TOPEX/Poseidon (T/P) satellite in 1992 and its suc- of the T/P SSH data allows us to not only document cessful multiyear measurements of the sea surface the low-frequency changes in a particular geographical height now provide a means to study the variability of domain, but also to test hypotheses and explore the un- gyre-scale ocean circulations. Using the T/P sea surface derlying dynamics of the observed circulation changes. height (SSH) data of the past eight years, the ®rst ob- Indeed, the major part of this study is devoted to pur- jective of this study is to quantify the large-scale chang- suing this second objective. es in the four current systems of the midlatitude sub- Being the out¯ow of an inertial western boundary tropical and subpolar North Paci®c. Note that the sub- current, the Kuroshio Extension is likely governed by polar North Paci®c is where large-amplitude high-fre- different dynamics than the NPC, the Alaska gyre, and quency SSH variability (wave periods shorter than a the WSG. In this study, we will focus on these latter few months) is detected by the T/P observations. The three current systems, whose large-scale changes are high-frequency SSH signals and their connections to the expected to be externally forced and governed by linear surface wind forcing have been the subject of several dynamics. Despite the expected linear dynamics, there

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC JANUARY 2002 QIU 355

FIG. 2. Distributions of the rms steric height variability due to (a) net surface heat ¯ux forcing and (b) evaporation minus precipitation (E Ϫ P) forcing. Units in centimeters. Calculated from the daily net surface heat ¯ux and E Ϫ P data of the NCEP reanalysis for the T/P period of Oct 1992±Jul 2000. remain questions and hypotheses that are not yet fully Paci®c Ocean. For example, are the SSH changes gen- answered and tested against observations. For example, erated in the eastern tropical Paci®c important for the does the SSH variability in a midlatitude subtropical variability in the Alaska gyre? In the western subarctic and subpolar ocean simply re¯ect changes in Ekman gyre, are the changes in the western boundary currents, ¯ux convergence/divergence? How important are bar- the EKC and the Oyashio, in balance with the time- oclinic adjustment processes for the SSH changes in a dependent interior Sverdrup ¯ow? Our goal is to address subpolar ocean? Given the slow propagation of mid- these questions under a simple and uni®ed dynamic latitude baroclinic Rossby waves, what role does eddy framework. dissipation play in determining the midlatitude SSH sig- Several recent studies have examined long-term sea nals? In addition to these questions that are general to surface temperature (SST) data in the Paci®c Ocean the SSH changes in any midlatitude ocean, there are (e.g., Miller et al. 1994; Deser and Blackmon 1995; also questions that are speci®c to the midlatitude North Nakamura et al. 1997). A common feature revealed by

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC 356 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32 these analyses is that the dominant mode of the SST ronmental Prediction (NCEP) reanalysis (Kalnay et al. anomalies that has decadal timescales in the North Pa- 1996). Speci®cally, ci®c has its center of action located in the Kuroshio± (hЈ ␣(tץ -Oyashio extension region (i.e., along the boundary be s [Q(t) Q(t)] (t)S (t)[F(t) F(t)], (1) tween the areas KE and WSG de®ned in Fig. 1b). As ϭ Ϫ Ϫ ␤ o Ϫ t ␳opcץ will become clear from this study, this is the region where signi®cant, low-frequency ocean circulation where ␳o, cp, ␣, ␤, and So are the reference density, the changes are detected. Clarifying the dynamics govern- speci®c heat, the thermal expansion rate, the saline con- ing the observed low-frequency SSH signals in the mid- traction rate, and the surface mixed layer salinity of latitude North Paci®c is, thus, also a needed step in seawater, respectively. In Eq. (1), Q(t) denotes the net improving our understanding of the ocean's role in de- surface heat ¯ux, F(t) the EϪ P rate, and overbars the termining the observed midlatitude SST anomalies. temporal average. To ®nd ␣(t), ␤(t), and So(t), we used This paper is organized as follows. Section 2 gives the monthly mixed layer temperature and salinity data a brief description of T/P SSH data processing with an of Levitus and Boyer (1994) and Levitus et al. (1994). emphasis on the removal of steric height signals. In In Figs. 2a and 2b, we plot the rms steric height section 3, we introduce a measure that quanti®es the variability associated with (a) the net surface heat ¯ux large-scale circulation ¯uctuations and describe the forcing and (b) the E Ϫ P forcing, respectively. Notice overall changes observed in the midlatitude North Pa- that the thermosteric height changes are dominated by ci®c Ocean. With the aid of a two-layer dynamic model, annual variations and their rms amplitudes range from detailed examinations into the causes of the observed 2.0 to 7.0 cm (Fig. 2a). Due to the dominance of the changes in the North Paci®c Current, the Alaska gyre, annual variations, the spatial pattern of the thermosteric and the western subarctic gyre are presented in sections height signals re¯ects that of the seasonally varying net 4, 5, and 6, respectively. Section 7, summarizes the surface heat ¯uxes. Compared to the thermosteric height results from this study. signals, the halosteric height signals have, in general, much smaller amplitudes (Fig. 2b). Aside from a lo- 2. T/P altimeter data calized area centered around 36ЊN, 150ЊW where the amplitude exceeds 0.6 cm, the overall rms amplitude of TOPEX/Poseidon altimeter data from the period of the salinehЈ signals is less than 0.4 cm in the region October 1992 through July 2000 (repeat cycles 2±287) s north of 35ЊN. That the thermosteric height changes are used in this study. For all the ground tracks passing dominate the total steric height signals is also found by the North Paci®c Ocean, we ®rst adjust the raw altimeter Gill and Niiler (1973) and Vivier et al. (1999). data for various environmental corrections, including the dry troposphere, the tides, and the inverse barometer It is worth pointing out that over the shallow conti- effect, according to the GDR users handbook (Callahan nental shelf regions along the Alaska/Canada coast, 1993). After these corrections, the height data are in- halosteric height signals have been observed to surpass terpolated to a common latitude grid with a one-per- the thermosteric height signals as a result of coastal second sample rate (about 5.6 km along a ground track). runoff (Royer 1979, 1981; Tabata et al. 1986). The dom- The alongtrack anomalous SSH pro®les are then cal- inance of the saline signals, however, diminishes off- culated by subtracting from each height pro®le the tem- shore of the continental slope (Tabata et al. 1986). Be- porally averaged pro®le (namely, the geoid plus the 8- cause the T/P SSH measurements are not adequately yr mean SSH). Finally, a low-pass ®lter, which has a accurate over the shallow shelf regions, our analysis is half-power point at 26 km, is applied to the anomalous con®ned to the offshore, open oceans (with water depth SSH data to suppress small-scale noise. Ͼ 1 km; see the shaded areas in Fig. 2), where the As recognized by several recent studies, a signi®cant thermosteric height signals dominate (hЈs ). part of the altimetrically measured SSH anomaly signals The strong spatial dependence of (hЈs ), shown in Fig. in the midlatitude oceans is due to the seasonally vary- 2a, indicates that the steric height signals cannot be ing surface heat ¯ux and evaporation minus precipita- eliminated by simply removing the basin-averaged SSH tion (E Ϫ P) forcing that causes expansion or contrac- anomalies from the individual T/P cycles (e.g., Isoguchi tion of the water column (e.g., Wang and Koblinsky et al. 1997). By so doing, one would introduce spatially 1996; Stammer 1997; Gilson et al. 1998; Vivier et al. varying, spurious annual variations to the resultant SSH 1999). These so-called steric height anomalies are large- signals, rendering dynamic interpretations of these sig- ly con®ned to the seasonal thermocline of the upper nals questionable. For data analyses throughout this ocean. Because the steric height change does not alter study, the SSH anomalies (hЈ) are de®ned as the T/P- the mass of seawater, it is dynamically passive and is measured SSH anomalies minus the thermal plus saline not of interest to this study. To remove the steric height steric height anomalies. In addition, hЈ signals with pe- signals (hЈs ), we estimate their time series along the T/ riods shorter than 70 days are removed in order to focus P ground tracks by using the daily net surface heat ¯ux on the large-scale, low-frequency variability of the and E Ϫ P data from the National Centers for Envi- ocean circulation.

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC JANUARY 2002 QIU 357

FIG. 3. ``Relative intensity'' of (a) the Western Subarctic Gyre, (b) the Alaska Gyre, (c) the Western Subarctic Gyre, and (d) the Kuroshio Extension. Here, the intensity is computed by ®rst projecting the residual geostrophic ¯ows (inferred from the T/P data) onto the Levitus climatological surface geostrophic streamlines, and then by averaging the projected residual geostrophic ¯ows in the speci®ed geographic area (see Fig. 1b).

3. Changes in the midlatitude North Paci®c changes will be pursued in the ensuing sections. To current systems quantify the large-scale changes of a current system measured by the T/P altimeters, we adopt in this section In this section, we describe the overall changes in the the following measure. Assuming geostrophy, the anom- four current systems in the midlatitude North Paci®c, alous surface geostrophic velocity is related to hЈ by namely, the western subarctic gyre, the Alaska gyre, the g (١hЈ, (2 North Paci®c Current, and the Kuroshio Extension. uЈϭ k ϫ More detailed analyses examining the causes for these g f

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC 358 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

FIG. 4. First EOF mode of the T/P-measured SSH anomalies in the subtropical and subpolar North Paci®c: the spatial pattern (top) and the corresponding weighting function (bottom). Units in meters. The steric height signals were removed prior to the EOF analysis. This mode explains 14.6% of the total variance.

where k is a unit vector in the vertical direction, f is case with the WSG during the study period, the pro- the Coliolis parameter, and g is the gravity constant. Let jections of the anomalies on the mean do not re¯ect the true changes in intensity, and the usefulness of VЈ(t) g de®ned by Eq. (4) diminishes (see section 6). (١h (3 ug ϭ k ϫ f Figure 3 shows the time series of VЈ for the four be the mean surface geostrophic velocity withh cal- current systems in the midlatitude North Paci®c (see culated from the Levitus climatology (Fig. 1a). Then, Fig. 1b for the geographical areas de®ning the current the area-averaged projection of the anomalous surface systems). Both the WSG and the Alaska gyre in Figs. geostrophic ¯ow along the mean surface geostrophic 3a and 3b show clear annual variations with a maximum streamlines, in January±March and a minimum in August±October. On average, the seasonal, peak-to-peak VЈ difference is Ϫ1 1 ug about 0.8 cm s for both the WSG and the Alaska gyre. VЈ(t) ϭ uЈg(t)´ dx dy, (4) With an average radius of R ϭ 1000 km for the WSG A ͵͵ |ug | A (R ϭ 700 km for the Alaska gyre) and assuming that provides one measure for the current intensity changes the anomalous surface geostrophic ¯ow extends to a in the area. In Eq. (4), A denotes the area of averaging. characteristic depth of Ho ϭ 1000 m (Ho ϭ 500 m), This measure, which we will term ``relative intensity,'' ⌬VЈϭ0.8 cm sϪ1 translates to a seasonal, peak-to-peak is appropriate when circulation anomalies are aligned volume transport change RHo⌬VЈ of ϳ8.0 Sv for the with the mean circulation, and proved to be a good WSG and ϳ2.8 Sv for the Alaska gyre, respectively (Sv measure of the true relative intensities of the NPC, the ϵ 106 m3 sϪ1). These transport numbers clearly depend

Alaska gyre, and the Kuroshio Extension. However, on the chosen values of R and Ho; nevertheless, the 2.8 when a gyre undergoes structural changes, as was the Sv value for the Alaska gyre is close to the observed,

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC JANUARY 2002 QIU 359

FIG. 5. SSH difference between the two triads: 1997/98/99 and 1993/94/95. Units in meters. Solid contours indicate positive SSH values and dashed lines, negative values.

peak-to-peak transport changes of the Alaskan Stream These interannual changes are shown by Qiu (1995, (2.4 Sv: Royer 1981). The 8.0 Sv value for the WSG 2000) to re¯ect the oscillation between the Kuroshio is smaller than the seasonal transport changes found in Extension's elongated mode and its contracted mode. In the Oyashio southeast of Hokkaido (ϳ20 Sv: Kono and the elongated mode, the Kuroshio Extension has a larger Kawasaki 1997). This could be due to the chosen char- eastward surface transport (which corresponds to a acteristic depth (Ho ϭ 1000 m) being too shallow for greater VЈ value in Fig. 3d), a more northerly axis po- the WSG. sition, and is associated with a more intense southern Figures 3a and 3b also reveal interannual changes in recirculation gyre. The reverse is true when the Kuro- the WSG and the Alaska gyre. With the exception of shio Extension is in the contracted mode. In comparison 1993 and 1998, the Alaska gyre shows a weak, but with the Kuroshio Extension system, our understanding generally increasing, trend in its intensity. The WSG, of the dynamics controlling the low-frequency vari- on the other hand, appears to have a different interannual ability in the WSG, the Alaska gyre, and the NPC is signal: it was relatively strong in 1993±96, weakened more limited. In the following sections, we will focus in 1997, and remained in a weakened state after 1998. our attention on these three current systems individually. As will become clear in section 6, this low-frequency signal only signi®es the interannual changes of the 4. The North Paci®c Current southern limb of the WSG (i.e., the Subarctic Current) and does not represent the intensity of the entire WSG. The North Paci®c Current variability is characterized The interannual changes are most clearly seen in the by a persistent interannual trend. It is instructive here North Paci®c Current (Fig. 3c). The steady intensi®- to examine this trend from a different perspective. To cation of the NPC from 1993 to 1998 has a ⌬VЈ of about do so, we conducted an empirical orthogonal function 1cmsϪ1. With a mean width of 1000 km and assuming (EOF) analysis of the SSH anomaly data in the region

Ho ϭ 500 m, this ⌬VЈ value translates to a 5 Sv volume north of 20ЊN. The ®rst EOF mode explains 14.6% of transport increase in the NPC. The intensifying trend of the total variance, and its spatial pattern and weighting the NPC appears to be leveling off following 1998. coef®cient are shown in Fig. 4. The weighting coef®- Notice that the NPC has a much weaker annual variation cient of this mode is dominated by a linear increasing than the two subpolar gyres we noted above. trend with seasonal modulations embedded. In the NPC Interannual variability of the Kuroshio Extension ex- region of the eastern North Paci®c (32Њ±45ЊN), the sign hibits a quite different characteristic (Fig. 3d). The Ku- of the spatial pattern is negative to the north and positive roshio Extension weakened gradually from late 1992 to to the south, suggesting that the zonal-mean meridional 1995±96 and reversed the weakening trend after 1997. SSH gradient across the NPC increased steadily over

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC 360 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32 the T/P period. The robustness of the spatial pattern of this EOF mode can be veri®ed if we simply take the 2 ϫ ␶ h1 fo١ץ f o ץ SSH difference between the three years in the beginning , 2h Ϫ (h Ϫ h ) ϩ ␤ ϭٌ x ␳ gHץtgH112ץ of the T/P mission, 1993/94/95, and the more recent []1 o 1 three years of 1997/98/99. The SSH pattern from this simple subtraction (see Fig. 5) resembles very much the (9) hץh ␤gЈץ gЈ f 2 ץ spatial pattern of the ®rst EOF mode and signi®es again (22h ϩٌh Ϫ o h ϩ ␤ 12ϩ ϭ 0, (10ٌ xץxgץtggH122ץ the enhancement of the north±south SSH gradient across []2 the NPC. To quantify the interannual changes in the NPC, we where ␤ is the meridional gradient of f and f o is the plot in Figs. 6a and 6b the time series of the SSH anom- Coriolis parameter at the reference latitude. alies averaged in a box north and south of the NPC It is useful here to express the SSH anomaly as (hЈ andhЈ ), respectively. Taking their difference and NS h1bcbtϵ h ϩ h , (11) multiplying it by Ϫg/ fW, where W is the meridional distance between the two boxes, gives a measure for the where hbc denotes that part of the SSH change due to zonal-mean surface ¯ow of the NPC. The NPC's relative the baroclinic processes, and hbt the part due to the bar- otropic processes. Equations governing h and h can intensity thus estimated (Fig. 6c) is very similar to that bc bt of Fig. 3c based on Eq. (4). Given the zonality of the be derived from Eqs. (9) and (10) as follows: ϫ ␶ hfH١ץ f 2 ץ NPC, this similarity is expected. There are two points (worth noticing from Fig. 6. First, bothhЈ andhЈ have ٌ2h Ϫ o h ϩ ␤ bc ϭ oe , (12 NS tgbcH bc x gH2 ␳ 1 ץ΂΃Ј eo ץ similar annual variations. This explains the weak annual ϫ ␶ hf١ץץ signals in the NPC. Second, the interannual trend in the 2 bt o (NPC's intensity is mainly due to the gradual SSH drop (ٌ hbt) ϩ ␤ ϭ , (13 ( x ␳ g(H ϩ Hץ tץ on the northern side of the NPC. As such, understanding o 12 the interannual changes of the NPC requires clarifying where He ϭ H1H2/(H1 ϩ H 2) denotes the equivalent the causes that determine the SSH signals on the north- depth in the two-layer model. Given that baroclinic ern side of the NPC. Rossby radii of deformation in the region of our interest As a dynamic framework to examine the SSH vari- are shorter than 40 km (e.g., Chelton et al. 1998), the ations, we adopt in this study a two-layer ocean model. long-wave approximation is justi®ed. Under this ap- Though simple in its formulation, the two-layer model proximation, Eq. (12) simpli®es to captures the barotropic and ®rst-mode baroclinic Rossby 2 ϫ ␶ ١ hgbc ЈHeץ hbc ␤gЈHeץ wave processes that are important for the large-scale ϪϭϪ. (14) x ␳ gf Hץtf22ץ SSH changes. In the two-layer model, linearized mo- ooo1 2 mentum and continuity equations are: Denoting cR ϭϪ␤gЈHe/f o , the solution to Eq. (14) can be found by integration along the long baroclinic Rossby u1 ␶ץ :ϩ f k ϫ u11ϭϪg١h ϩ , (5) wave characteristic t ␳o H1ץ x Ϫ x h (x, t) ϭ hx, t Ϫ e ץ u ϭ 0, (6) bc bc e ´ h12Ϫ h ) ϩ H 11١) t ΂΃cRץ

x u2ץ fHoe xϪ xЈ ,ϫ ␶ xЈ, t Ϫ dxЈ ١ ϩ f k ϫ u212ϭϪg١h Ϫ gЈ١h , (7) Ϫ t 2ץ ␳o g␤Hc1 ͵ R x ΂΃ e (h2 (15ץ (u ϭ 0, (8 ´ ١ ϩ H t 22ץ where xe is the longitudinal location of the eastern boundary. Barotropic ocean responses to ¯uctuating sur- where ui is the ith-layer velocity vector, Hi is the mean face winds are timescale dependent. For timescales lon- thickness of the ith-layer, h1 is the SSH anomaly, h 2 is ger than months, Willebrand et al. (1980) show that the the layer interface anomaly, ␶ is the surface wind stress time derivative term in the barotropic vorticity equation is negligible and that Eq. (13) in this case reduces to vector, ␳o is the reference density, and gЈϭ(␳ 2 Ϫ ␳1)g/ the time-dependent Sverdrup balance: ␳o is the reduced gravity. For simplicity, we have taken H ϭ const in Eq. (8). Effects of bottom topography ϫ ␶ hfbt o١ץ 2 will be discussed toward the end of this section. As- ϭ . (16) ( x ␳ g␤(H ϩ Hץ suming that the timescales of interest are long compared o 12 with the inertial period, we can combine Eqs. (5)±(8) Estimating hbt and hbc from Eqs. (16) and (15) requires ϫ ␶ ®eld. In this study, we use the monthly wind ١ to form the following vorticity equations: the

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC JANUARY 2002 QIU 361 . Dashed line: SSH anomalies measured by the Ј bt h ϩ ЈЈ bt bc hh . 7. (a) Barotropic SSH anomalies, , calculated from the time-dependent Sverdrup IG , calculated from Eq. (15). Dashed line: SSH anomalies measured by the T/P altimeters; F Ј bc balance [Eq. (16)] inh the box northsame of as the in NPC. Fig. (b)T/P 6a. Solid altimeters. (c) line: Solid baroclinic SSH line: anomalies, . / Њ g Ϫ and a meridional dimension of 5 Њ . 1 Ϫ s 4 Ϫ 10 ϫ 0.91 ϭ f lat and Њ 8 ϭ W . 6. SSH anomalies averaged in (a) a box north of the NPC and (b) a box south of where IG F fW, (c) The SSH anomaly difference between the north and south boxes. SSH is scaled by the NPC. Both boxes have a zonal dimension of 30

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC 362 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

FIG. 8. SSH anomalies due to anomalous Ekman pumping [Eq. (17)] in the box north of the NPC. stress dataset from the NCEP reanalysis (Kalnay et al. Several investigators in the past have hypothesized 1996). The horizontal resolution of the NCEP data is that the SSH anomalies may simply re¯ect convergence/ 1.904Њ lat ϫ 1.825Њ long. In Fig. 7a, we plot the time divergence of the anomalous surface Ekman ¯uxes (e.g., series of the hbt anomalies in the region north of the Kelly et al. 1993; Bhaskaran et al. 1993; Lagerloef NPC (40Њ±45ЊN, 160Њ±130ЊW) computed from Eq. (16). 1995). In other words, one may simplify the baroclinic In accordance with the theory of Willebrand et al. vorticity equation (14) to (1980), wind stress curl signals shorter than two months ϫ ␶ hgЈH2١ץ -have been removed. Also, H ϩ H ϭ 5000 m is as 1 2 Ek ϭϪ e . (17) t ␳ gf H2ץ sumed. Figure 7a shows that the barotropic SSH changes in this region are commonly small, O(2 cm), except oo1 during wintertime when the Aleutian low intensi®es and To test this hypothesis, we plot in Fig. 8 the time series extends eastward over the subpolar North Paci®c Ocean of the hEk anomalies in the region north of the NPC (Rienecker and Ehret 1988; Strub and James 2000a, using Eq. (17) and the parameter values speci®ed above. manuscript submitted to Progress in Oceanography, AlthoughhEkЈ exhibits large interannual changes, it lacks hereafter SJa). Although the hbt anomalies show signif- the interannual trends that characterize thehЈ time series icant winter-to-winter changes, the overall interannual bc (Fig. 7b). In general, a difference betweenhEkЈ and h bcЈ signal inhbtЈ is of a different nature from that detected implies that the wind stress curl forcing in the past and in the T/P data (Fig. 6a). over the region to the east of the domain of interest is To estimate the baroclinic SSH anomalies using Eq. important. Since the domain of our interest is located (15), we adopt the following parameter values: gЈϭ close to the eastern boundary, we expect that the steady 0.03 m sϪ2 (White 1982) and H ϩ H ϭ 5000 m. For 1 2 SSH drop, which is predicted inhЈ for the T/P period the long baroclinic Rossby wave speed c , we use the bc R but not inh , is due to the wind stress curl forcing prior latitude-dependent values presented in Killworth et al. EkЈ (1997). Along 42.5ЊN, for example, c ϭϪ1.3 cm sϪ1. to October 1992. To verify this point, we plot in Fig. 9 ϫ ␶ anomalies along 41.9ЊN as a function of time ١ R the By requiring the long baroclinic Rossby wave speed, cR ϭϪ␤gЈH H f 2/(H ϩ H ) to match the theoretical val- and longitude. The straight line in the ®gure indicates 1 2 o 1 2 the phase speed of the long baroclinic Rossby wave at ue of Killworth et al., we obtain H1 ϭ 264 m and He Ϫ1 ϭ 250 m for the region north of the NPC. The solid this latitude: cR ϭϪ1.43 cm s . Notice that the wind stress curl anomalies are persistently negative during line in Fig. 7b shows the time series of the hbc anomalies based on these parameter values. Compared to the T/P- the period from 1986 to 1990. This long-lasting negative wind stress curl results in Ekman ¯ux convergence and derived SSH signals (hЈN , the dashed line), the baroclinic SSH anomalies capture well the two observed inter- is the reason why the observed SSH anomalies at the annual trends, namely, the steady SSH drop from late beginning of the T/P mission were relatively high. From 1992 to 1998 and the reversing trend following 1998. 1991 to 1998, Fig. 9 shows that the wind stress curl 1 The predictive skill ofhЈN byhЈbc in Fig. 7b is S ϭ 0.50. anomalies north of the NPC were largely positive. It is ϫ ␶ЈϽ0 in 1986±90 to ١ Notice that while adding the barotropic SSH signals to this decadal transition from

ϫ ␶ЈϾ0 in 1991±98 that causes the decreasing trend ١ hbcЈ improves the comparison with the observed SSH signals on the annual timescale (see Fig. 7c), it is the inhbcЈ shown in Fig. 7b. Notice that the wind stress curl baroclinic SSH changes that determine the interannual anomalies after 1998 change the sign to negative; this signal inhЈN . transition from a positive to a negative wind stress curl forcing is responsible for the reversal in the SSH trend 1 Here, predictive skill is de®ned by S ϭ 1 Ϫ͗(h Ϫ h )2͘/͗͘h2 , seen inhЈN andhbcЈ following 1998. o p o The lower layer thickness H of our two-layer model where ho is the observed signal, hp is its prediction, and angle brackets 2 denote time averaging. has been assumed to be constant. In the limit of H1/(H1

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC JANUARY 2002 QIU 363

FIG. 9. Wind stress curl anomalies along 41.9ЊN as a function of time and longitude. Units in 10 Ϫ8 NmϪ1. For clarity, signals with periods shorter than 6 months have been removed. The straight thin line denotes the Ϫ1 characteristic of long baroclinic Rossby waves at this latitude (cR ϭϪ1.43 cm s ). The SSH signal at a point on this line is determined by the wind stress curl forcing accumulated along the characteristic to the east of the point.

2 ϩ H 2) K 1, a variable bottom topography will have the barotropic vorticity equation. Speci®cally, the time- little effect upon the baroclinic SSH changes governed dependent Sverdrup balance is now by Eq. (14). Allowing a variable H 2, however, does alter ϫ ␶ ff١ Jh, ϭ , (18) bt 2 ΂΃H12ϩ H ␳o g(H12ϩ H ) 2 Typical depth ratio for the midlatitude North Paci®c Ocean is ϳ0.05. where J is the Jacobian operator. Instead of latitudinal

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC 364 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

lines, hbt in Eq. (18) can again be evaluated by integra- whereas the baroclinic SSH signals capture the inter- tion along constant f/(H1 ϩ H 2) contours emanating annual trend present in thehЈO time series (the dashed from the eastern boundary. Notice that the maps pre- line in Fig. 11b). Examining the wind stress curl chang- sented by Koblinsky et al. (1989, Plate 2) and Vivier es along 53.3ЊN across the center of the Alaska gyre et al. (1999, their Fig. 10) show that, except very close (Fig. 12) reveals that the decreasing trend inhbcЈ is due to the eastern boundary, large-scale f/(H1 ϩ H 2) con- to the fact that negative wind stress curl anomalies over tours in the NPC region do not deviate substantially the Alaska gyre in 1989±93 produced positive SSH from the latitudinal lines. Given the large spatial scales anomalies in 1993, whereas the wind stress curl anom- of the surface wind forcing, it is reasonable to expect alies were largely positive following 1994. This again that inclusion of the bottom topography in our two-layer demonstrates the importance of the wind stress curl model will not signi®cantly change thehbtЈ time series forcing prior to the launch of the T/P satellite in con- presented in Fig. 7a. tributing to the SSH trends detected during the T/P For the region south of the NPC, solving Eq. (15) period. indicates that the baroclinic SSH variability there has Figure 11c shows the comparison betweenhЈO and the an interannual amplitude much smaller than its coun- sum of the barotropic and baroclinic SSH signals. The terpart to the north of the NPC. Because the interannual high predictive skill ofhЈO by hЈbtϩ h bcЈ (S ϭ 0.84) in- changes in the NPC are largely due to the SSH changes dicates that much of the observed, large-scale SSH sig- on the northern side of the NPC (Fig. 6), detailed results nals in the offshore Alaska gyre can be explained by onhЈS will not be presented here. In concluding this the barotropic and ®rst-mode baroclinic Rossby wave section, we again emphasize the importance of the plan- dynamics. This favorable comparison reinforces the etary vorticity advection term in Eq. (14). The baroclinic point we raised in section 3 of appropriately removing SSH signals are not determined by the Ekman ¯ux con- the steric height signals prior to analyzing the T/P SSH vergence or divergence; instead, they represent the in- data. The dashed line in Fig. 13a shows the T/P SSH tegrated outcome of the past wind stress curl forcing anomalies in the offshore Alaska gyre region without along baroclinic Rossby wave characteristics. the removal of the steric height signal. Even though the steric height signal in this region is comparatively small (recall Fig. 2), failing to properly subtract it would nev- 5. The Alaska gyre ertheless obscure the conclusion that the two-layer dy- In the Alaska gyre region north of 48ЊN and east of namics form an adequate basis for understanding the 165ЊW, the ®rst EOF mode of the observed SSH anom- observed large-scale SSH anomalies. alies shows that the center of the Alaska gyre has On the annual timescale, Fig. 10c shows that the shoaled steadily over the past eight years (Fig. 4). From intensity of the Alaska gyre peaks in winter. This ®nd- geostrophy, this implies a strengthening trend in the ing is consistent with the studies by Bhaskaran et al. cyclonic circulation of the Alaska gyre. The result pre- (1993) using the Geosat altimeter data, and by SJa sented in Fig. 3b, on the other hand, indicates that the using the T/P altimeter data. As a possible cause for intensity of the Alaska gyre has a more complicated the winter maximum of the gyre intensity, Bhaskaran interannual signal than a well-de®ned strengthening et al. suggested seasonal Ekman pumping, although no trend. To clarify this apparent discrepancy, we plot in quantitative argument was provided in their study. To Fig. 10 the SSH anomalies in a 2Њ band along the Alaska/ assess the effect of the Ekman pumping forcing, we

Canada coast from 130Њ to 165ЊW (hЈC ) and the SSH plot in Fig. 13b the Ekman pumping±induced SSH anomalies in the Alaska gyre offshore of this coastal anomalies calculated from Eq. (17). The seasonal peak- band (hЈO ). Consistent with the steady shoaling signal to-peak amplitude inhEkЈ is ϳ2 cm, which is less than found in the EOF analysis, thehЈO time series in Fig. half of the observed amplitude of 5 cm (Fig. 10a). 10a shows a clear decreasing trend. The SSH anomalies Based on the result shown in Fig. 11, we believe that in the coastal band reveal an interannual signal of a the annual modulation in the Alaska gyre intensity considerably different nature (Fig. 10b):hЈC was high in should be interpreted as the barotropic response to the 1992/93 and in late 1997/early 1998, and dropped to a seasonal wind stress curl forcing. The magnitude of persistent low level after mid-1998. Not surprisingly, it this response is quanti®able by the time-dependent is the difference betweenhЈCO andhЈ that determines the Sverdrup balance. intensity of the Alaska gyre (cf. Fig. 10c with Fig. 3b). For the interannual changes in the Alaskan gyre in- In the following, we will try to clarify the forcing mech- tensity, Fig. 10 shows clearly the contribution from the anisms underlying thehЈOC andhЈ signals. SSH signal along the Alaska/Canada coast. That the For the offshore Alaska gyre region, we plot in Figs. alongshore SSH anomalies were high during late 1997/ 11a and 11b the barotropic and baroclinic SSH anom- early 1998 points to their connection to the tropical, alies calculated from Eqs. (16) and (15), respectively. ENSO-related SSH variability. Indeed, plotting the SSH The results are in general similar to those discussed anomalies along the eastern boundary of the North Pa- above for the NPC; namely, the barotropic SSH signals ci®c from the equator to 60ЊN and along the Alaska are dominated by the annual and intraannual variations, coast from 142Њ to 165ЊW (Fig. 14) reveals the general

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC JANUARY 2002 QIU 365 N. The dashed Њ offshore from the Alaska/ Њ W and to the south along 50 Њ . 11. Same as in Fig. 7 except for the region that is 2 IG F lines in (b) andFig. (c) 10a. show the SSH anomalies measured by the T/P altimeters; same as in Canada coast, bounded to the west along 165 Њ to 8 N. Њ Њ ϭ W offshore from where Њ fW, / g Ϫ W and to the south along 50 Њ OC ЈЈ hh band along the Alaska/Canada coast from 165 Њ . 1 Ϫ s 4 Ϫ 10 ϫ 1.18 ϭ f . 10. (a) SSH anomalies averaged in the region that is more than 2 W. (c) SSH anomaly difference between and scaled by IG Њ F 130 (b) SSH anomalies averaged in a 2 the Alaska/Canada coast, bounded to the west along 165 lat and

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC 366 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

point between stations to the south where sea level is in¯uenced by anomalies of equatorial origin and stations to the north where sea level is in¯uenced by local wind stress anomalies.'' Interestingly, a close look at Fig. 14 shows some evidence that the T/P SSH anomalies north and south of 38ЊN do have subtly different coherent alongshore structures. In order to quantify the relative importance of locally and remotely forced SSH changes along the Alaska/ Canada coast, we consider here the dynamic balance proposed by En®eld and Allen (1980) and Chelton and Davis (1982): hЈ ␶Јץ 1 ϭ s , (19) s ␳o gH1ץ

where s denotes the alongshore distance and␶Јs , the alongshore component of the wind stress anomaly vec- tor. Equation (19) can be derived from Eq. (5) by as- suming the timescales of interest are long compared with those of the coastal trapped waves. Physically, it rep-

resents the balance between the␶Јs -induced onshore Ek- man transport (if ␶Јs Ͼ 0) and the offshore geostrophic transport associated with the alongshore SSH anomaly gradient. For the SSH anomaly averaged along a coast

segment from se to sw, integrating Eq. (19) leads to

sssww 11␶ sЈ(sЈ) h11Ј ds ϭ hЈ(se) ϩ dsЈ ds, (20) LL͵ ͵͵␳o gH1 ssseee

where L is the length of the coast segment andh1Ј (se) FIG. 12. Wind stress curl anomalies along 53.3ЊN as a function of is the SSH anomaly at the upstream boundary, s ϭ se. time and longitude. Units in 10Ϫ8 NmϪ1. For clarity, signals with In contrast to this remote upstream signal, the second periods shorter than 6 months have been removed. The straight thin line denotes the characteristic of long baroclinic Rossby waves at term on the rhs of Eq. (20) denotes the SSH anomalies Ϫ1 this latitude (cR ϭϪ0.80 cm s ). The SSH signal at a point on this forced by the local alongshore wind ¯uctuations. line is determined by the wind stress curl forcing accumulated along In Fig. 15, we compare the observed SSH anomalies the characteristic to the east of the point. averaged along the Alaska/Canada coast from 130Њ to 165ЊW with (a) the observed SSH anomalies at the

upstream location se : 130ЊW and 50ЊN, (b) the local correspondence between the SSH anomaly signals along alongshore wind-forced SSH anomalies [i.e., the sec- the Alaska/Canada coast and those in the Tropics. This ond term on the rhs of Eq. (20)], and (c) the sum of alongshore correspondence in SSH signals has been the upstream and locally forced SSH anomalies. The found by many previous investigators based on tide predictive skills ofhЈC by (a), (b), and (c) are S ϭ gauge sea level data analyses (e.g., En®eld and Allen Ϫ0.12, Ϫ1.94, and 0.35, respectively. Clearly, it is the 1980; Chelton and Davis 1982; Emery and Hamilton combination of the local and upstream forcing that de- 1985). It is also noticed by Strub and James (2000b, termines the T/P measured SSH anomalies along the manuscript submitted to Progress in Oceanography)in Alaska/Canada coast. The importance of this combined their analysis of the T/P data with a focus on 1997±98 forcing, which was also emphasized in the study by El NinÄo anomalies. While recognizing the impact of the En®eld and Allen (1980), is well illustrated by the high equator-originated SSH anomalies upon those along the SSH anomaly event that appeared from November midlatitude coast through coastal trapped wave dynam- 1997 to March 1998. After mid-1997, Fig. 15a shows ics, the aforementioned investigators have also empha- that the SSH anomalies upstream of the Alaska/Canada sized the importance of the alongshore SSH changes coast had two maxima with the peaks around July 1997 resulting from local surface wind forcing. The relative and February 1998, respectively. This double-maxi- importance of these two coexisting forcing mechanisms mum signal is ``transformed'' into a single-maximum appears to depend on the geographical distance from signal at the Alaska/Canada coast owing to the locally the equator. For example, in their analysis of the long- forced SSH anomalies that have a dominant winter- term tide gauge sea level data, En®eld and Allen (1980) maximum and summer-minimum annual cycle (Fig. noted that ``San Fransisco (at 37.8ЊN) may be a dividing 15b). As shown in Fig. 15c, the single-maximum signal

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC JANUARY 2002 QIU 367

ЈЈ FIG. 13. (a) Solid line: hhbtϩ bc in the offshore Alaska Gyre (same as the solid line in Fig. 11c). Dashed line: T/P SSH anomalies without removal of the steric height signal. (b) SSH anomalies due to anomalous Ekman pumping [Eq. (17)] in the offshore Alaska Gyre.

thus combined matches favorably with the observed the ``relative intensity'' in 1993±95 is due to the inten-

hЈC signal. si®cation of the eastward ¯owing Subarctic Current Following the 1997±98 El NinÄo event, the SSH along whenhЈNS was in the low state andhЈ , the high state. the Alaska/Canada coast drops and remains at a low WhenhЈN becomes high after 1997, the reduced Subarctic level. Figure 15b shows that this is in part due to the Current contributes to the weakening of the ``relative smaller SSH changes forced by the local alongshore intensity'' of the WSG. winds during the winters of the succeeding La NinÄa Because the large-scale SSH anomaly pattern in the years. That the Aleutian low over the Alaska gyre tends WSG is not aligned with the Levitus mean SSH pattern, to weaken following an El NinÄo event is consistent with caution is needed in interpreting the time series of Fig. the ®nding by Kelly et al. (1993), who showed that the 3a. By adding the SSH anomalies of 1993±95 and Aleutian low was weaker in the winters of 1987/88 and 1997±99 respectively to the Levitus mean SSH ®eld 1988/89 as compared to the preceding winter when the (Figs. 17a and 17b) we ®nd that the WSG underwent 1986±87 El NinÄo was in the mature phase. a structural change between these two periods. The WSG was more zonally elongated and accompanied by an intenser Subarctic Current in 1993±95. In contrast, 6. The western subarctic gyre the WSG was more zonally contracted in 1997±99. The EOF analysis of the T/P SSH anomalies (Fig. 4) Although the Subarctic Current was weaker during this shows that the large-scale SSH signals in the WSG re- latter period, Fig. 17b shows that the western limb of gion are coherent in three subregions: a northern sub- the WSG (the East Kamchatka Current and the Oyash- region centered around 47ЊN and 178ЊE wherein the io) was stronger than in 1993±95. The WSG's struc-

SSH anomalies (hЈN ) had an increasing trend over the T/ tural change, of course, results from the large-scale P period, and southern and western subregions wherein SSH changes in the three subregions of the WSG. Un-

the SSH anomalies (hЈSW andhЈ ) decreased generally over derstanding this structural change thus requires clari- the T/P period. As shown in Fig. 16a, the time series fying these SSH signals.

ofhЈN is characterized by a steplike transition from a In order to understand thehЈN signals, we plot in Figs. relatively low SSH state (1993±95) to a relatively high 18a and 18b the barotropic and baroclinic SSH anom-

SSH state (1997±99). In comparison to thehЈN signals, alies calculated from Eqs. (16) and (15). Similar to the the interannual trends inhЈSW andhЈ are opposite and are results of the preceding sections, thehbtЈ signals are dom- generally less prominent (Figs. 16b and 16c). Much of inated by the annual timescale ¯uctuations, whereas the

the interannual variability we found in the ``relative in- hbcЈ signals capture the SSH changes on interannual time- tensity'' of the WSG (Fig. 3a) is associated with this scales. Although the comparison between the observed

steplike change inhЈN . Speci®cally, the strengthening of hЈN time series and that of hbtЈ ϩ hЈ bc appears reasonable

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC 368 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

FIG. 14. T/P SSH anomalies averaged within the 1Њ band along the Paci®c Ocean eastern boundary from the equator to 60ЊN and along the Alaska coast from 142Њ to 165ЊW as a function of time. Contour units in 0.05 m.

(Fig. 18c), the predictive skill ofhЈN by hbtЈ ϩ h bcЈ in this Physically,hbcЈ is the result of the wind stress curl case is rather low: S ϭϪ0.56. As is clear in Fig. 18b, forcing accumulated over the journey a baroclinic Ross- the reason for this low skill is that the baroclinic SSH by wave takes from the eastern boundary along the signal based on Eq. (15) overestimates the interannual North America coast to the subregion of interest. Be- change inhЈN :hЈbc was estimated too low in 1993/94 and cause of the slow propagation of the baroclinic Rossby too high following 1997. waves at the WSG latitude, the time span of the journey

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC JANUARY 2002 QIU 369 . 16. SSH anomalies averaged in (a) the northern, (b) the southern, and (c) the western IG F subregions of theare Western determined Subarctic from the Gyre. spatial pattern The(Fig. of the 4). geographical ®rst EOF locations mode of of the the observed SSH subregions anomalies band Њ N, (b) the local alongshore wind forced SSH anomalies, Њ W (dashed lines) with (a) the observed SSH anomalies at the Њ W and 50 Њ to 165 Њ . 15. Comparisons between the T/P measured SSH anomalies along the Alaska/Canada IG F coast from 130 upstream location, 130 and (c) the sumthe of assumption the underlying upstream Eq. and (19), locally all forced quantities SSH shown anomalies. here To are be averaged consistent in with the 1 along the coast.

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC 370 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

FIG. 17. Sea surface height (mean ϩ anomalies) ®eld averaged in (a) 1993±95 and (b) 1997±99. Here, the mean SSH ®eld is the Levitus climatology (Fig. 1a). Shaded areas have the SSH values Ͻ1.1 m.

ϫ ␶ hgЈH2١ץ h ␤gЈHץ ,is on the order of 15 years. Over such a long journey bc ϪϭϪϪe bc e ⑀h , (21) x ␳ gf H bcץtf22ץ it is highly likely that the SSH signals generated in the past will be partially dissipated. In order to quantify ooo1 where ⑀ denotes the Newtonian dissipation rate. Inte- how eddy dissipation affects the baroclinic SSH chang- grating along the baroclinic Rossby wave characteristic, es, we add a Newtonian dissipation term to Eq. (14): the solution to Eq. (21) is

x Ϫ x ⑀ fHx xϪ xЈ ⑀ (ϫ ␶ xЈ, t Ϫ exp Ϫ (x Ϫ xЈ) dxЈ. (22 ١ h (x, t) ϭ hx, t Ϫ eoeexp Ϫ (x Ϫ x ) Ϫ bc bc ee2 ccRR␳ og␤Hcc1 ͵ RR ΂΃x ΂΃ []e []

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC JANUARY 2002 QIU 371

Ј FIG. 18. (a) Barotropic SSH anomalieshbt in the northern subregion of the WSG calculated Ј from Eq. (16). (b) Solid line: baroclinic SSH anomalieshbc calculated from Eq. (15). Dashed line: SSH anomalies measured by the T/P altimeters; same as in Fig. 16a. (c) Solid line: ЈЈ hhbtϩ bc. Dashed line: SSH anomalies measured by the T/P altimeters.

Since the dissipation rate is a poorly known quantity, quires the consideration of eddy dissipation. Using a we calculate thehbcЈ signals using a wide range of the forced baroclinic Rossby wave model, Schneider et al. ⑀ values. Figure 19a shows the predictive skill ofhЈN by showed the dissipation rate to be 1/(4 yr), a rate quite hbtЈ ϩ h bcЈ as a function of ⑀. Inclusion of the dissipative similar to the optimal ⑀ value found in this study. effect tends to reduce the magnitude of thehbcЈ signals Along the Kuril Islands ¯ow the East Kamchatka Cur- and signi®cantly improve the predictive skill. From Fig. rent (EKC) and its southern continuation, the Oyashio. 19a, a maximum skill, S ϭ 0.49, is found when ⑀Ϫ1 ϭ Many previous studies have considered the EKC/Oyashio 6 yr (compare with S ϭϪ0.56 for ⑀ ϭ 0). The com- variability to be constrained by the time-dependent Sver- parison betweenhЈN and the hbtЈ ϩ hЈ bc time series cal- drup balance. The validity of the Sverdrup balance has culated using this optimal ⑀ value is shown in Fig. 19b. been suggested from indicative subsurface water tem- It is worth noting that the optimal dissipation rate thus perature data (Sekine 1988; Hanawa 1995), seasonal hy- estimated is small. This suggests that while eddy dis- drographic observations (Kono and Kawasaki 1997), and sipation plays an important role for the baroclinic SSH the SSH-derived Oyashio transport based on the ®rst 2- signals in the WSG, it is less important in the NPC and yr T/P data (Isoguchi et al. 1997). While it is well es- the Alaska gyre regions; baroclinic Rossby waves can tablished that the EKC/Oyashio transport is in phase with traverse these regions from the eastern boundary within the Sverdrup transport on the annual timescale, the ques- the 6-yr e-folding timescale of eddy dissipation. By an- tion of whether the time-dependent Sverdrup balance also alyzing results from a coupled ocean±atmosphere GCM, explains the interannual changes in the EKC/Oyashio is Schneider et al. (2002) found that explaining the ther- unanswered. As we have found throughout this study, mocline changes in the Kuroshio±Oyashio extension re- baroclinic processes are important for the large-scale in- gion as a result of basin-scale surface wind forcing re- terannual SSH signals. As such, we cannot take for grant-

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC 372 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

FIG. 19. (a) Predictive skill of the observed SSH anomalies in the northern subregion of the ЈЈЈ Ј WSG (hhhN ) by btϩ bcas a function of the dissipation rate ⑀. Here, h bc is calculated from Eq. ЈЈЈϪ1 (22). (b) Comparison betweenhhhN (dashed line) and btϩ bc (solid line) when ⑀ ϭ 6yr. ed that the time-dependent Sverdrup balance should be velocity. The Sverdrup signal appears also to capture valid beyond the annual timescale. some of the year-to-year changes in the wintertime EKC/ To address this question, we ®rst quantify the vari- Oyashio velocity peaks. For example, both time series ability of the EKC/Oyashio. Figures 20a and 20b show show that the velocity peaks in 1993, 1995, 1996, and the SSH anomalies averaged in the two 2Њ-wide bands 1998 are larger than those in 1994, 1997, and 1999. bracketing the EKC/Oyashio from 43Њ to 51ЊN. From Notwithstanding these agreements, Fig. 21b reveals that the difference between these two time series, we plot a lower-frequency difference exists between the two in Fig. 20c the surface velocity anomalies of the EKC/ time series. As indicated by the solid line in Fig. 21c, Oyashio. Because of the abundance of mesoscale eddies the scaled Sverdrup velocity is consistently larger than offshore of the Kuril Islands (Talley 1991; Rogachev the observed EKC/Oyashio velocity in 1992±95, where- 2000; Yasuda et al. 2000), we are using the along- as it is consistently smaller in 1996±98. boundary averaged velocity anomalies to represent the Could this low-frequency difference in the velocity be EKC/Oyashio variability. Notice that the offshore band explained by the baroclinic ocean responses that are ig-

(hЈO ) has, in general, larger and better de®ned annual- nored in the time-dependent Sverdrup balance? To clarify to-interannual SSH changes than its coastal counterpart this, we calculate the baroclinic SSH anomalies,hbcЈ , in (hЈC ) and that the EKC/Oyashio velocity anomaly is the offshore band of the Kuril Islands from Eq. (22). Ϫ1 largely determined byhЈO . (From the result shown in Fig. 19, ⑀ ϭ 6 yr is again In Fig. 21a, we plot the Sverdrup transport anomalies used.) The dashed line in Fig. 21c shows the surface averaged from 43Њ to 51ЊN by integrating the NCEP EKC/Oyashio velocity changes induced by the offshore wind stress curl data from the eastern boundary to off- baroclinic SSH anomalies (i.e., ϪghbcЈ / fW). The time se- shore of the Kuril Islands. To compare this time series ries has a low-frequency signal very similar to that of with the SSH-derived EKC/Oyashio velocity time se- the solid line (the linear correlation coef®cient between ries, we scale the Sverdrup transport value by Ϫ0.0026 the two time series in Fig. 21c is 0.69). Speci®cally, in msϪ1 SvϪ1 and superimpose the two time series in Fig. 1993±95 the baroclinic Rossby wave responses increase 21b. Here, the scaling factor is determined by maxi- the SSH offshore of the Kuril Islands and work to weaken mizing the predictive skill of the SSH-derived EKC/ the surface EKC/Oyashio velocity, whereas in 1996±98 Oyashio velocity by the Sverdrup velocity. As shown the SSH anomalies offshore of the Kuril Islands is low- in Fig. 21b, the annual cycle of the EKC/Oyashio ve- ered by the baroclinic Rossby wave processes, giving locity agrees very well with that of the scaled Sverdrup rise to a stronger EKC/Oyashio velocity. The result of

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC JANUARY 2002 QIU 373 . 1 Ϫ Sv 1 Ϫ 0.0026 m s Ϫ fW. / Ј bc h g Ϫ N. (b) Solid line: Sverdrup transport anomalies scaled by Њ ±51 . 21. (a) Sverdrup transport anomalies averaged along offshore of the Kuril Islands Њ IG F of 43 Dashed line: EKC/Oyashio velocityin anomalies Fig. estimated 20c. fromOyashio (c) the Solid velocity SSH gradient; line: anomalies. samebaroclinic Difference Dashed as SSH between line: changes the offshore EKC/Oyashio Sverdrup of velocity and the anomalies the Kuril induced SSH-derived Islands, EKC/ i.e., by Њ Њ 2 ϭ W where fW, / g Ϫ band along the Kuril Islands from 43 Њ band offshore of the coastal band de®ned Њ OC ЈЈ hh C Ј h . 1 O Ј Ϫ h s 4 Ϫ 10 ϫ 1.07 ϭ f . 20. (a) SSHN. anomalies (b) ( SSH anomalies ( ) averaged in ) a averaged in 2 a 2 Њ IG F to 51 in (a). (c) SSHlong and anomaly difference between and scaled by

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC 374 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 32

Fig. 21 suggests that while the barotropic Sverdrup re- de®ned annual cycle: a maximum in January±March and sponses are responsible for the EKC/Oyashio variability a minimum in July±September. This annual cycle is a on the annual and year-to-year timescales, contributions consequence of the barotropic Sverdrup response of the from the baroclinic Rossby wave responses are important Alaska gyre to the seasonally varying surface wind on the multiyear timescales. stress curl forcing. The intensity of the western subarctic gyre also un- dergoes a noticeable annual modulation with a winter 7. Summary maximum and a summer minimum. Like the Alaska Altimetric observations from the TOPEX/Poseidon gyre, the amplitude and phase of the modulation are satellite over the past eight years provide us an un- determined by the time-dependent barotropic Sverdrup precedented sea surface height dataset to investigate the balance. Interannually, the WSG transformed from a ocean circulation changes on various timescales. Using zonally elongated gyre in 1993±95 to a zonally more this dataset, we examined the large-scale circulation contracted gyre in 1997±99. The elongated WGS had changes in the midlatitude North Paci®c Ocean. Partic- the more intense eastward ¯owing Subarctic Current, ular attention was paid to the North Paci®c Current, the whereas the contracted WSG, with the gyre center shift- Alaska gyre, and the western subarctic gyre. In addition ed westward, had the more intense East Kamchatka Cur- to the description of the large-scale changes in these rent and Oyashio. This structural change of the WSG current systems, the goal of this study was to construct is due to the steady SSH rise in the northern subregion a dynamic framework that can aid our understanding of of the WSG and the steady SSH drop in its southern the observed circulation changes on the seasonal-to-in- and western subregions. Although the baroclinic Rossby terannual timescales. To achieve this goal, we adopted wave dynamics are again important for these interannual a two-layer ocean model that takes into account the ®rst- SSH changes, we found that to adequately simulate the mode baroclinic Rossby wave dynamics and the time- observed SSH trends requires taking into account the dependent barotropic Sverdrup dynamics. Though sim- eddy dissipation effect. By maximizing predictive skill, ple in dynamics, we found that this model is adequate we estimated the timescale for the Newtonian dissipa- to reproduce many of the observed large-scale SSH sig- tion rate to be about six years. Dynamically, this implies nals. The baroclinic Rossby wave dynamics explain the that the wind stress curl forcing over the eastern Alaska interannual SSH changes consistently in all of the three gyre exerts less of an in¯uence upon the interannual current systems of our interest. Similarly, the barotropic changes of the WSG than does the forcing over the Sverdrup dynamics are responsible for the observed western Alaska gyre. For the western boundary currents SSH signals on the annual and shorter timescales. in the subpolar North Paci®c, we found that the EKC/ The large-scale variability in the North Paci®c Cur- Oyashio variability is in balance with the interior Sver- rent is dominated by its steady strengthening over the drup ¯ow variability on the annual and year-to-year period from late 1992 to 1998. Much of this interannual timescales. On the multiyear timescales, the SSH signals signal is due to the large-scale SSH dropping on the governed by the baroclinic Rossby wave dynamics are northern side of the NPC. With the slow propagation of important in explaining the low-frequency modulation the baroclinic Rossby waves at the NPC latitude and of the EKC/Oyashio changes that is unaccounted for by with its presence near the eastern boundary, the inter- the barotropic Sverdrup balance. annual signal in the NPC at a particular time depends on the regional wind stress curl forcing accumulated Acknowledgments. This study bene®ted from many over the years prior to that time. Because the large-scale discussions with Ted Durland, John Hargrove, Humio SSH signals throughout the NPC region have similar Mitsudera, and Ted Strub. Detailed comments made by annual variations, the intensity of the NPC itself exhibits Lynne Talley and an anonymous reviewer helped clarify no clear modulation on the annual timescale. many parts of an early version of the manuscript. The In the Alaska gyre interior, the SSH anomalies show surface wind stress and heat ¯ux data were provided by a similar interannually decreasing trend as in the north- the National Center for Environmental Prediction and ern NPC region. In other words, the Alaska gyre interior the TOPEX/Poseidon data, by the Physical Oceanog- shoaled steadily over the period from 1993 to 1999. The raphy DAAC at Jet Propulsion Laboratory. I am grateful intensity of the Alaska gyre, however, depends not only to Gary Mitchum for his help in processing the original on the interior SSH anomalies, but also on those along T/P data. Support from NASA through the TOPEX/ the Alaska/Canada coast. We found that the SSH anom- Poseidon Extended Mission (Contract 960889) and the alies along the Alaska/Canada coast are determined by Jason1 project (Contract 1207881) is gratefully ac- a combination of remotely and locally forced signals. knowledged. The alongshore SSH was anomalously high during the 1997±98 El NinÄo event (Aug 1997±Mar 1998) and REFERENCES dropped to a persistently low level immediately follow- Bhaskaran, S., G. S. E. Lagerloef, G. H. Born, W. J. Emery, and R. ing the El NinÄo event. In addition to the interannual R. Leben, 1993: Variability in the Gulf of Alaska from Geosat signal, the intensity of the Alaska gyre also has a well- altimetry data. J. Geophys. Res., 98, 16 311±16 330.

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC JANUARY 2002 QIU 375

Callahan, P. S., 1993: TOPEX/POSEIDON GDR user handbook. Jet Miller, A. J., D. R. Cayan, T. P. Barnett, N. E. Graham, and J. M. Propulsion Laboratory Special Publication D-8944, 84 pp. Oberhuber, 1994: Interdecadal variability of the Paci®c Ocean: Chelton, D. B., and R. E. Davis, 1982: Monthly mean sea-level var- Model response to observed heat ¯ux and wind stress anomalies. iability along the west coast of North America. J. Phys. Ocean- Climate Dyn., 9, 287±302. ogr., 12, 757±784. Mizuno, K., and W. B. White, 1983: Annual and interannual vari- ÐÐ, R. A. DeSzoeke, M. G. Schlax, K. E. Naggar, and N. Siwertz, ability in the system. J. Phys. Oceanogr., 13, 1998: Geographical variability of the ®rst baroclinic Rossby ra- 1847±1867. dius of deformation. J. Phys. Oceanogr., 28, 433±460. Musgrave, D. L., T. J. Weingartner, and T. C. Royer, 1992: Circulation Deser, C., and M. L. Blackmon, 1995: On the relationship between and hydrography in the northwestern Gulf of Alaska. Deep-Sea tropical and North Paci®c sea surface variations. J. Climate, 8, Res., 39, 1499±1519. 1677±1680. Nakamura, H., G. Lin, and T. Yamagata, 1997: Decadal climate var- Emery, W. J., and K. Hamilton, 1985: Atmospheric forcing of inter- iability in the North Paci®c during the recent decades. Bull. annual variability in the northeast Paci®c Ocean. J. Geophys. Amer. Meteor. Soc., 78, 2215±2225. Res., 90, 857±868. Qiu, B., 1995: Variability and energetics of the Kuroshio Extension En®eld, D. B., and J. S. Allen, 1980: On the structure and dynamics and its recirculation gyre from the ®rst two-year TOPEX data. of monthly mean sea level anomalies along the Paci®c coast of J. Phys. Oceanogr., 25, 1827±1842. North and South America. J. Phys. Oceanogr., 10, 557±578. ÐÐ, 2000: Interannual variability of the Kuroshio Extension and its Favorite, F., A. J. Dodimead, and K. Nasu, 1976: Oceanography of impact on the wintertime SST ®eld. J. Phys. Oceanogr., 30, the subarctic Paci®c region, 1962±72. Bull. Int. North Pac. Fish. 1486±1502. Comm., 33, 1±187. Reed, R. K., 1984: Flow of the Alaskan Stream and its variations. Fu, L.-L., and R. A. Davidson, 1995: A note on the barotropic re- Deep-Sea Res., 31, 369±386. sponse of sea level to time-dependent wind forcing. J. Geophys. Rienecker, M. M., and L. L. Ehret, 1988: Wind stress curl variability Res., 100, 24 955±24 963. over the North Paci®c from the Comprehensive Ocean±Atmo- Fukumori, I., R. Raghunath, and L.-L. Fu, 1998: Nature of global sphere Data Set. J. Geophys. Res., 93, 5069±5077. large-scale sea level variability in relation to atmospheric forc- Rogachev, K. A., 2000: Recent variability in the Paci®c western ing: A model study. J. Geophys. Res., 103, 5493±5512. subarctic boundary currents and Sea of Okhotsk. Progress in Gill, A., and P. P. Niiler, 1973: The theory of the seasonal variability Oceanography, Vol. 47, Pergamon, 209±336. in the ocean. Deep-Sea Res., 20, 141±177. Royer, T. C., 1979: On the effect of precipitation and runoff on coastal Gilson, J., D. Roemmich, B. Cornuelle, and L.-L. Fu, 1998: Rela- circulation in the Gulf of Alaska. J. Phys. Oceanogr., 9, 555± tionship of TOPEX/Poseidon altimetric height to steric height 563. and circulation in the North Paci®c. J. Geophys. Res., 103, ÐÐ, 1981: Baroclinic transport in the Gulf of Alaska. Part I. Sea- 27 947±27 965. sonal variations of the Alaska Current. J. Mar. Res., 39, 239± Hanawa, K., 1995: Southward penetration of the Oyashio water sys- 250. tem and the wintertime condition of midlatitude westerlies over Schneider, N., A. J. Miller, and D. W. Pierce, 2002: Anatomy of North the North Paci®c. Bull. Hokkaido Natl. Fish. Res. Inst., 59, 103± Paci®c decadal variability. J. Climate, in press. 119. Sekine, Y., 1988: Anomalous southward intrusion of the Oyashio east Isoguchi, O., H. Kawamura, and T. Kono, 1997: A study on wind- of Japan. 1. In¯uence of the seasonal and interannual variations driven circulation in the subarctic North Paci®c using TOPEX/ in the wind stress over the North Paci®c. J. Geophys. Res., 93, POSEIDON altimeter data. J. Geophys. Res., 102, 12 457±12 468. 2247±2255. Kalnay, E., and Coauthors, 1996: The NCEP/NCAR 40-Year Re- Stammer, D., 1997: Steric and wind-induced changes in TOPEX/ POSEIDON large-scale sea surface topography observations. J. analysis Project. Bull. Amer. Meteor. Soc., 77, 437±471. Geophys. Res., 102, 20 987±21 009. Kawai, H., 1972: Hydrography of the Kuroshio Extension. Kuro- Tabata, S., 1991: Annual and interannual variability of baroclinic shioÐIts Physical Aspects, H. Stommel and K. Yoshida, Eds., transports across Line P in the northeast Paci®c Ocean. Deep- University of Tokyo Press, 235±354. Sea Res., 38, S221±S245. Kelly, K. A., M. J. Caruso, and J. A. Austin, 1993: Wind-forced ÐÐ, B. Thomas, and D. Ramsden, 1986: Annual and interannual variations in sea surface height in the northeast Paci®c Ocean. variability of steric sea level along Line P in the northeast Paci®c J. Phys. Oceanogr., 23, 2392±2411. Ocean. J. Phys. Oceanogr., 16, 1378±1398. Killworth, P. D., D. B. Chelton, and R. A. De Szoeke, 1997: The Talley, L. D., 1991: An Okhotsk sea water anomaly: Implications for speed of observed and theoretical long extratropical planetary ventilation in the North Paci®c. Deep-Sea Res., 38, S171±S190. waves. J. Phys. Oceanogr., 27, 1946±1966. Vivier, F., K. A. Kelly, and L. Thompson, 1999: The contributions Koblinsky, C. J., P.P.Niiler, and W. J. Schmitz Jr., 1989: Observations of wind forcing, waves, and surface heating to sea surface of wind-forced deep ocean currents in the North Paci®c. J. Geo- height observations in the Paci®c Ocean. J. Geophys. Res., 104, phys. Res., 94, 10 773±10 790. 20 767±20 788. Kono, T., and Y. Kawasaki, 1997: Results of CTD and mooring ob- Wang, L., and C. J. Koblinsky, 1996: Annual variability of the sub- servations southeast of Hokkaido. 1. Annual velocity and trans- tropical recirculations in the North Atlantic and North Paci®c: port variations in the Oyashio. Bull. Hokkaido Natl. Fish. Res. A TOPEX/POSEIDON study. J. Phys. Oceanogr., 26, 2462± Inst., 61, 65±81. 2479. Lagerloef, G., 1995: Interdecadal variations in the Alaska Gyre. J. White, W. B., 1982: Traveling wave-like mesoscale perturbations in Phys. Oceanogr., 25, 2242±2258. the North Paci®c Current. J. Phys. Oceanogr., 12, 231±243. Levitus, S., and T. Boyer, 1994: World Ocean Atlas 1994. Vol 4: Willebrand, J., S. G. H. Philander, and R. C. Pacanowski, 1980: The Temperatures, NOAA Atlas NESDIS 4, U.S. Govt Printing Of- oceanic response to large-scale atmospheric disturbances. J. ®ce, Washington, 117 pp. Phys. Oceanogr., 10, 411±429. ÐÐ, R. Burgett, and T. Boyer, 1994: World Ocean Atlas 1994. Vol Yasuda, I., and Coauthors, 2000: Cold-core anticyclonic eddies south 3: Salinity, NOAA Atlas NESDIS 3, U.S. Govt Printing Of®ce, of the Bussol' Strait in the northwestern Subarctic Paci®c. J. Washington, 99 pp. Phys. Oceanogr., 30, 1137±1157.

Unauthenticated | Downloaded 09/26/21 12:27 PM UTC