Why Is the Westward Rossby Wave Propagation from the California Coast “Too Fast”?

Allan J. Clarke aDepartment of Earth, Ocean, and Atmospheric Science, Florida State University, Tallahassee, Florida

Search for other papers by Allan J. Clarke in
Current site
Google Scholar
PubMed
Close
and
Sean Buchanan aDepartment of Earth, Ocean, and Atmospheric Science, Florida State University, Tallahassee, Florida

Search for other papers by Sean Buchanan in
Current site
Google Scholar
PubMed
Close
Free access

Abstract

Past work has shown that interannual California coastal sea level variability is mostly of equatorial origin, and decades of satellite sea surface height (SSH) and in situ dynamic height observations indicate that this interannual signal propagates westward from the California coast as nondispersive Rossby waves (RWs). These observations agree with standard linear vertical mode theory except that even when mean flow and bottom topography are considered, the fastest baroclinic vertical mode RW in each case is always much slower (1.6–2.3 cm s−1) than the observed 4.2 cm s−1. This order-1 disagreement is only resolved if the standard bottom boundary condition that the vertical velocity w′ = 0 is replaced by perturbation pressure p′ = 0. Zero p′ is an appropriate bottom boundary condition because south of San Francisco the northeastern Pacific Ocean boundary acts approximately like an impermeable vertical wall to the interannual equatorial wave signal, and therefore equatorial quasigeostrophic p′ is horizontally constant along the boundary. Thus, if equatorial p′ = 0 at the bottom, then this condition also applies off California. The large-scale equatorial ocean boundary signal is due to wind-forced eastward group velocity equatorial Kelvin waves, which at interannual and lower frequencies propagate at such a shallow angle to the horizontal that none of the baroclinic equatorial Kelvin wave signal reaches the ocean floor before striking the eastern Pacific boundary. Off California this signal can thus be approximated by a first baroclinic mode with p′ = 0 at the bottom, and hence the long RW speed there agrees with that observed (both approximately 4.2 cm s−1).

Significance Statement

The California Current System is one of the most biologically rich and best-documented coastal regions in the world. In this region coastal sea level propagates westward from the coast at about 110 km month−1, slow enough to enable us to make large-scale ocean climate forecasts of the California Current ecosystem using coastal sea level. Although the westward speed seems slow, theoretically it is about double what we would expect. Offered here is an explanation of why this speed is “too fast” by linking the California wave signal to the equator, El Niño, and the shallow equatorial ocean response to the wind.

© 2024 American Meteorological Society. This published article is licensed under the terms of the default AMS reuse license. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Allan J. Clarke, aclarke@fsu.edu

Abstract

Past work has shown that interannual California coastal sea level variability is mostly of equatorial origin, and decades of satellite sea surface height (SSH) and in situ dynamic height observations indicate that this interannual signal propagates westward from the California coast as nondispersive Rossby waves (RWs). These observations agree with standard linear vertical mode theory except that even when mean flow and bottom topography are considered, the fastest baroclinic vertical mode RW in each case is always much slower (1.6–2.3 cm s−1) than the observed 4.2 cm s−1. This order-1 disagreement is only resolved if the standard bottom boundary condition that the vertical velocity w′ = 0 is replaced by perturbation pressure p′ = 0. Zero p′ is an appropriate bottom boundary condition because south of San Francisco the northeastern Pacific Ocean boundary acts approximately like an impermeable vertical wall to the interannual equatorial wave signal, and therefore equatorial quasigeostrophic p′ is horizontally constant along the boundary. Thus, if equatorial p′ = 0 at the bottom, then this condition also applies off California. The large-scale equatorial ocean boundary signal is due to wind-forced eastward group velocity equatorial Kelvin waves, which at interannual and lower frequencies propagate at such a shallow angle to the horizontal that none of the baroclinic equatorial Kelvin wave signal reaches the ocean floor before striking the eastern Pacific boundary. Off California this signal can thus be approximated by a first baroclinic mode with p′ = 0 at the bottom, and hence the long RW speed there agrees with that observed (both approximately 4.2 cm s−1).

Significance Statement

The California Current System is one of the most biologically rich and best-documented coastal regions in the world. In this region coastal sea level propagates westward from the coast at about 110 km month−1, slow enough to enable us to make large-scale ocean climate forecasts of the California Current ecosystem using coastal sea level. Although the westward speed seems slow, theoretically it is about double what we would expect. Offered here is an explanation of why this speed is “too fast” by linking the California wave signal to the equator, El Niño, and the shallow equatorial ocean response to the wind.

© 2024 American Meteorological Society. This published article is licensed under the terms of the default AMS reuse license. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Allan J. Clarke, aclarke@fsu.edu

1. Introduction

In agreement with basic linear large-scale, low-frequency dynamics, Clarke and Dottori (2008, hereafter CD08; see also Dottori and Clarke 2009), showed that observed coastal sea level and dynamic height and satellite sea surface height (SSH) propagate westward from the California coast like nondispersive Rossby waves (RWs). This interannual and decadal variability in sea level and dynamic height is mostly of equatorial origin via the ocean coastal waveguide (Enfield and Allen 1980; Chelton and Davis 1982; Clarke and Lebedev 1999; Kessler 1990; Frischknecht et al. 2015), and huge interannual and decadal fluctuations in the California Current ecosystem are intimately connected with the equatorial signal (Chelton et al. 1982; Clarke and Lebedev 1999; CD08; Dottori and Clarke 2009; Frischknecht et al. 2015). Based on the RW westward propagation physics, CD08 used the lead time of the monthly coastal sea level to show that monthly San Diego negative sea level anomalies could predict the log of the California zooplankton 2–4 months in advance with statistically significant correlations greater than 0.6. Remote equatorial forcing and RW dynamics is thus an essential mechanism in the prediction of the large-scale California Current physics and ecosystem.

Although the independent SSH and California Cooperative Oceanic Fisheries Investigation (CalCOFI) dynamic height and coastal sea level observations seemed to agree well with the theory in that westward long RW propagation was observed, at 4.2 cm s−1 the standard first vertical mode RW speed is about double that expected. Furthermore, CD08 found that even when the standard linear theories were modified to include zonal mean flow (Killworth et al. 1997) or coupled with topography (Killworth and Blundell 1999, 2003), the resultant wave speed estimates were still far too low (1.6–2.3 cm s−1). Tailleux and McWilliams (2001) suggested that if somehow the deep ocean could be “decoupled” from the upper ocean, then the RW speed could be adjusted to agree with observations, but it was not clear to CD08 how this “bottom pressure decoupling,” or p′ = 0 bottom boundary condition could be justified. Tailleux and McWilliams (2001) proposed that standard constant-depth flat-bottom vertical modes under the influence of bottom topography or a mean flow may decorrelate and give rise to zero bottom pressure. However, LaCasce and Groeskamp (2020) pointed out that bottom pressure decoupling seems conceptually problematic, since it depends on the fast barotropic mode interacting with a much slower first baroclinic flat bottom mode. In addition, although a barotropic mode and a first baroclinic mode can be added together so that their bottom pressures add to zero, the resulting structure does not agree with observations as well as the vertical structure of a first baroclinic mode with p′ = 0 replacing vanishing vertical velocity on the ocean floor (see, e.g., Figs. 2 and 3 in de La Lama et al. 2016).

Negligible horizontal flow and p′ = 0 at the ocean floor rather than vanishing vertical velocity are consistent with observations in other far more general contexts than California (Wunsch 1997; Wortham and Wunsch 2014; de La Lama et al. 2016). Various explanations for negligible flow and bottom pressure have been offered, including distortion of the pressure and horizontal flow by the mean flow (Killworth et al. 1997; Brink and Pedlosky 2020), dissipation of the large-scale flow by bottom friction (Brink and Pedlosky 2018), modification of the bottom boundary condition by large-scale bottom slope (Killworth and Blundell 1999; LaCasce 2017), or rough bottom topography (Samelson 1992; Bobrovich and Reznik 1999; LaCasce and Groeskamp 2020). In theory rough bottom topography with bottom slope greater than about 10−3 causes the large scale flow to vanish and p′ = 0 at the bottom, and global estimates of bottom slope for topography at 1-min horizontal resolution (see Fig. 12 of de La Lama et al. 2016) show that indeed small scale bottom slope does exceed 10−3 for almost all of the ocean floor, including the region of interest off California. Furthermore, both Tailleux and McWilliams (2001) and LaCasce and Groeskamp (2020) show that the p′ = 0 bottom boundary condition does lead to Rossby wave phase speed increases comparable to the observed increases.

However, most of the low-frequency variability in the ocean is forced by the wind at the sea surface, and the variability seen at the bottom does not reach the bottom instantly or locally. In that case the bottom pressure perturbation and associated flow generated by the wind would be zero at the bottom and p′ = 0 would be an appropriate bottom boundary condition. We suggest that this is the case for the interannual Rossby wave signal off California. Specifically, we present theory showing that the wind-driven interannual equatorial El Niño ocean signal does not reach the ocean floor by the time it reaches the eastern boundary, and that therefore there is no remote wind-driven energy at the deep ocean floor. Consequently, p′ = 0 at the eastern boundary ocean floor at the equator. At these low frequencies the flow is quasigeostrophic, and, since there is no flow into the eastern boundary, p′ is constant spatially along the boundary and so vanishes at the bottom at California.

Tailleux and McWilliams (2001) and LaCasce and Groeskamp (2020) showed that globally the p′ = 0 bottom boundary condition results in increased long RW speeds, and in the deep sea off California the speeds are approximately in agreement with the observed CD08 speeds. For the case of constant depth water, long Rossby waves propagate westward at a speed proportional to (f)−2, which implies that over the California Current region stretching from about 27° to 39°N the long RW speed increases southward by a factor of 1.9. In addition, in the southern part of the California Current System (CCS) the bottom topography is shallower near the coast than in the northern part, and this may also affect the RW speed measured from the coast asymmetrically. To examine in more detail how the RW speed varies over the CCS region, we decided (section 2) to update the CD08 analysis with higher-resolution calculations of the locally constant depth RW speeds in both the w′ = 0 and p′ = 0 cases.

As has already been discussed and shown theoretically and observationally in the literature, the large-scale low-frequency signal south of about San Francisco is of equatorial origin (Enfield and Allen 1980; Chelton and Davis 1982; Clarke 1983). We briefly review this result with up-to-date satellite SSH data in section 3 and confirm that, within 100 km of the coast from 35°S to 35°N, the deep-sea interannual equatorial signal dominates the SSH and therefore the near-surface pressure signal. The equatorial theory and physical discussion in section 4 suggest that this low-frequency surface signal should propagate downward at such a shallow angle that there is no baroclinic bottom pressure at the eastern ocean boundary, and that therefore that p′ = 0 rather than w′ = 0 is a more efficient bottom boundary condition to use for the vertical modes in this case. In section 5, we discuss why the first baroclinic vertical mode should dominate the wind-driven ocean response. Section 6 briefly examines the applicability of our results to regions other than California, and concluding section 7 summarizes the main results.

2. Long Rossby wave speed estimates off California

To examine the influence of p′ = 0 on the bottom with higher resolution than previous analyses, including the effect of shallower water on the bottom boundary condition, we calculate theoretical long RW speeds in the same way as CD08 with updated data. Climatological values of u¯ and N2(z) for the years 1955–2012 at a 0.25° horizontal resolution were computed from the World Ocean Atlas 2013, version 2 (WOA13-V2; Locarnini et al. 2013; Zweng et al. 2013), hydrographic dataset, and then used to compute the first baroclinic mode, local nondispersive long RW speeds. International Thermodynamic Equation Of Seawater—2010 (TEOS-10) thermodynamic subroutines (http://www.teos-10.org/) were used to compute N2 values from WOA13-V2. At interannual frequencies, the baroclinic velocity u¯ is in geostrophic balance and can be computed through vertically integrating meridional gradients in density via thermal wind balance. See Killworth et al. (1997) for more details regarding the eigenvalue problem that incorporates u¯.

Locally computed nondispersive long RW speeds for the first baroclinic mode are displayed in Fig. 1 within the CCS (27°–37°N, 115°–125°W) for depths greater than 3 km under the conditions (i) w′ = 0 at z = −H, (ii) w′ = 0 at z = −H with u¯ (Killworth et al. 1997), (iii) p′ = 0 at z = −H, and (iv) p′ = 0 at z = −H with u¯. Some general features of the plots in Fig. 1 are apparent. First, the RW speeds are generally much faster for the p′ = 0 bottom boundary condition (Figs. 1c,d) than the w′ = 0 bottom boundary condition (Figs. 1a,b). Second, in the standard no-mean-flow case, the long RW speed is βc12f2 where f is the Coriolis parameter, β is its northward gradient and c1 is the internal gravity wave speed for the first vertical mode. Since f−2 increases southward in the California Current as it is nearer to the equator, long RW speeds should increase southward in the no-mean-flow case, and Fig. 1 shows that this is a dominant tendency for all cases, not just the zero-mean-flow case. Figure 1 shows that the p′ = 0 bottom boundary condition gives RW speeds much closer to those observed, but in order to compare our results to CD08’s observed speeds quantitatively, we will need to calculate average theoretical wave speeds for the whole region.

Fig. 1.
Fig. 1.

Contour plot of the local first baroclinic mode Rossby wave speed in deep waters (>3 km) under the conditions (a) w′ = 0 at z = −H, (b) w′ = 0 at z = −H with u¯ (see Killworth et al. 1997), (c) p′ = 0 at z = −H, and (d) p′ = 0 at z = −H with u¯. The Rossby wave speeds (cm s−1) are positive for westward propagation.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0024.1

Average theoretical wave speeds for the California Current region were estimated in a similar way to CD08 by grouping and averaging theoretical lags in 130-km zonal distances found from local first vertical mode wave speeds, and then estimating the average speed via linear regression (see Fig. 2). Linear regressions of these lags and distances for either the w′ = 0 and p′ = 0 bottom boundary conditions give the estimated zonal theoretical speeds summarized in Table 1 and Fig. 2. Figure 2a displays the regression estimate of the longwave RW speed when lags for the first 260 km are excluded. When w′ = 0 is the bottom boundary condition, the estimated RW speed is 2.41 cm s−1 (green line) for u¯=0, but decreases to 2.13 cm s−1 with the inclusion of baroclinic zonal mean flow u¯ (blue line; Killworth et al. 1997). The corresponding results for the p′ = 0 bottom boundary condition are 4.12 cm s−1 when u¯=0 (red line) and 4.22 cm s−1 when u¯ is nonzero (black line). Both p′ = 0 speeds are much closer to the CD08 observed average RW speed at 4.2 cm s−1. Error bars are included for each theoretical wave speed estimate and are based on the standard deviation of the averaged lags (horizontal bars) and zonal distance covered in each group (vertical bars). Table 1 provides 95% confidence intervals on our wave speed estimates (slopes).

Fig. 2.
Fig. 2.

Theoretical Rossby wave speeds for the first baroclinic mode for (a) the exclusion (“Standard”) and (b) inclusion (“Shallower Lags Included”) of grouped lags for the first 260 km from the coast. In each panel, zonal distance is regressed on theoretical lags grouped by 130-km zonal distance increments away from the California coast based on either bottom boundary condition w′ = 0 and p′ = 0 with and without baroclinic zonal mean flow inclusion (Killworth et al. 1997). Vertical error bars encompass the 130-km zonal distance per increment, while horizontal error bars are a standard deviation of the lags used to average within each increment. Note: the 95% confidence intervals found in Table 1 capture the variability depicted by the error bars on the wave speed estimates.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0024.1

Table 1.

Summary of estimated westward-propagating nondispersive Rossby wave speeds (cm s−1) in the CCS (27°–37°N, 115°–125°W) for the first baroclinic mode for both the standard vanishing vertical velocity (w′ = 0) and vanishing baroclinic pressure (p′ = 0) bottom boundary conditions for both zero and nonzero mean zonal flow. Theoretical estimates were computed using the results shown in Fig. 1 and a linear regression analysis of grouped monthly lags in 130-km zonal distance increments from the coastline in the same manner as CD08 (Fig. 2). Observed wave speeds were obtained from hydrographic-derived dynamic height and satellite altimetry by CD08. The inclusion of a truncated u¯ and influence of shallower water lags on wave speed estimates (see text) are also reported. Included are 95% confidence intervals on the wave speed estimates in parentheses where applicable.

Table 1.

The choice to exclude theoretical lags within 260 km from the coast, as did CD08, is due to the prominence of shallow waters (<3 km) within 30°–35°N (Fig. 1). This region was included in CD08’s RW speeds estimated from the satellite altimetry and hydrographically derived dynamic height observations, but theoretical calculations with that region included only slightly changed the results. Specifically, under the p′ = 0 bottom boundary condition the estimated RW speed is 4.08 and 4.33 cm s−1 with and without u¯ inclusion, respectively; when w′ = 0 is used the RW speed is estimated as 2.20 and 2.54 cm s−1 (Fig. 2b). Thus, including the shallower region in the south does not change our conclusion that p′ = 0 is responsible for RW speeds observed by CD08. Additionally, due to hydrographic data limitations, CD08 set u¯=0 below 500 m. Figure 3 reveals that beneath 500 m the averaged u¯ profile in the CCS is much smaller than the representative wave speed of 4 cm s−1, so that we would not expect u¯ beneath 500 m to strongly influence the wave speed. In keeping with this, Table 1 shows that truncating u¯ to zero beneath 500 m changes the RW speeds only slightly, and it is still the change from the w′ = 0 to p′ = 0 bottom boundary condition that is most responsible for increasing the theoretical RW speed to near that observed.

Fig. 3.
Fig. 3.

Averaged zonal geostrophic velocity in the CCS (27°–37°N, 115°–125°W) based on WOA13-V2 climatological data for the years 1955–2012 for depths greater than or equal to 3 km.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0024.1

3. The eastern Pacific Ocean boundary signal at interannual and lower frequencies

a. California Rossby waves are predominately of equatorial origin

Figure 4 shows that interannual anomalies in coastal sea level at San Diego, California (32.7°N, 117.2°W), to be strongly correlated with the sea level at Santa Cruz Island, Ecuador (0.6°S, 90.3°W), and the Niño-3.4 SST index, respectively. Past work has also documented that interannual and lower-frequency coastal sea level signals from the equator are prominent along the eastern Pacific boundary in both Southern and Northern Hemispheres (Bretschneider and McLain 1976; Chelton 1980; Enfield and Allen 1980; Chelton and Davis 1982; Strub and James 2002).

Fig. 4.
Fig. 4.

Monthly time series for interannual anomalies for coastal sea level measurements at San Diego, California (blue line; 32.7°N, 117.2°W), Santa Cruz, Ecuador (green line; 0.6°S, 90.3°W), and interannual anomalies of the Niño-3.4 SST index (averaged over the box 5°S–5°N, 170°–120°W). Each time series is created by removing the seasonal cycle and filtered using an 11-point symmetric interannual low-pass filter (Trenberth 1984). Each time series is continuous from January 1970 to December 2016 except for the Santa Cruz station where infrequent sampling becomes commonplace after May 2003. The maximum correlation between every time series occurs at lag-zero and is r = 0.72 [rcrit(99%) = 0.48] between San Diego vs Santa Cruz sea levels, r = 0.75 [rcrit(99%) = 0.42] for San Diego sea level vs Niño-3.4 SSTs, and r = 0.74 [rcrit(99%) = 0.49] for Santa Cruz sea level vs Niño-3.4 SSTs. All critical correlation coefficients, here and elsewhere, are computed using Ebisuzaki’s (1997) nonparametric statistical significance test.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0024.1

At interannual frequencies the no normal flow boundary condition implies that the quasigeostrophic deep-sea pressure signal should be spatially constant and in phase along the boundary, so that an empirical orthogonal function (EOF) analysis of the SSH in the deep sea near the coast should result in complete dominance by the leading EOF whose spatial structure function would be constant along the deep-sea eastern boundary (Clarke and Shi 1991). We tested this using zonally averaged (1° longitude) JPL Making Earth Science Data Records for Use in Research Environments (MEaSUREs) SSH monthly anomalies off the eastern Pacific boundary from 35°S to 35°N (Fig. 5, top panel). Figure 5 shows that the first EOF does indeed dominate with 86% of the explained variance (Fig. 5, middle), but the structure function shows that the equatorial signal has an overall tendency to decrease poleward. This is due to several factors, including the increasing influence of other secondary signals like alongshore wind forcing (Enfield and Allen 1980), variations in coastline direction (Clarke and Shi 1991; Li and Clarke 2004), and dissipation of the large-scale signal due to eddies (Giunipero and Clarke 2013). Even though the EOF structure function is not spatially constant but decays poleward, it is clear from the leading EOF that the deep-sea signal near the surface and coast is dominated by the equatorial signal. Consistent with previous work, Fig. 5 (lower panel) shows that the principal component time series for the eastern boundary sea level signal is strongly correlated with the El Niño SST index Niño-3.4 {r = 0.81 [rcrit(99%) = 0.52]}. In other words, in agreement with theory and past analyses there is a strong, baroclinic interannual El Niño pressure signal at the surface that reaches California.

Fig. 5.
Fig. 5.

(top) Hovmöller time series of filtered (Trenberth 1984) interannual JPL MEaSUREs (Zlotnicki et al. 2019) SSH anomalies (cm) zonally averaged 1° off the eastern Pacific boundary for the latitudinal range 35°S–35°N. Our zonal average was taken to be 1° off the coast to capture the boundary signal. (middle) The primary EOF structural function of the zonally averaged interannual JPL SSH with percentage of explained variance (PEV = 86%). (bottom) Time series of the first principal component (blue line) and Niño-3.4 (orange line) with a maximum autocorrelation at lag 0 at r = 0.81 [rcrit(99%) = 0.52]. An interannual 11-point symmetric low-pass filter (Trenberth 1984) was utilized to capture the dominant interannual signal. This EOF analysis is an extension of Clarke (1992) who used 4 coastal stations from 15°S to 10°N to argue coastal pressure is spatially constant at low frequencies.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0024.1

b. Westward propagation of the boundary signal

Coastal interannual and lower-frequency sea level signals of equatorial origin have been observed to propagate offshore like long RWs as far south as Chile (Vega et al. 2003; Challenor et al. 2004; Ramos et al. 2008; Vergara et al. 2017) and as far north as California (Herrera Cervantes and Parés-Sierra 1994; Ramp et al. 1997; Fu and Qiu 2002; CD08; Frischknecht et al. 2015). Such a large-scale nondispersive westward-propagating signal from an eastern boundary in the form of RWs is in accordance with low-frequency theory (Cane and Sarachik 1977; McCreary 1977; Schopf et al. 1981; Cane and Moore 1981; Clarke 1983, 1992; Grimshaw and Allen 1988; Clarke and Shi 1991), which states that RWs are reflected wave solutions of incident longwave equatorial Kelvin waves (EKWs) at interannual frequencies. Additionally, it has been argued that local alongshore wind stress, especially in California (Clarke and Lebedev 1999; Clarke 1992; Dottori and Clarke 2009; Frischknecht et al. 2015) is a secondary contributor to the interannual coastal sea level signal compared to the boundary signal generated remotely at the equator.

Previously it has been assumed that this interannual pressure signal has the vertical structure of a standard first baroclinic mode with w′ = 0 as a deep-sea bottom boundary condition. But in section 2 we confirmed that the California RW propagation from the coast is far better described by a first vertical mode with p′ = 0 as the bottom boundary condition. To understand the vertical structure of this signal, we next consider its origin.

c. Physics of the interannual equatorial eastern boundary deep-sea pressure signal

The only way low-frequency, large-scale energy can reach the eastern equatorial Pacific boundary is via EKWs, since these are the only large-scale waves with eastward group velocity. Such waves are generated by low-frequency zonal equatorial wind stress surface forcing west of about 150°W (see Fig. 6), and so east of 150°W the eastward group velocity signal can be approximated as freely propagating EKWs. In fact, analysis shows (see, e.g., Clarke 1992) that the eastern boundary sea level is 2 times the EKW sea level at the equator. An unforced EKW is really a low-frequency internal gravity wave trapped by rotation to the equator, and, for buoyancy frequency N and wave frequency ω, the energy propagates downward at an angle ω/N to the horizontal. Such a vertically propagating signal can be described by the complete set of standard vertical modes with w′ = 0 on the ocean floor. Clarke and Thompson (1984) showed that for a similar problem of a wind-forced coastal Kelvin wave with an infinitesimally thin surface mixed layer the infinite set of modes can be summed analytically to give the downward propagation of alongshore current at angle ω/N to the horizontal. In a key equatorial study, McCreary (1984) obtained a similar result by numerically summing the standard vertical modes with w′ = 0 on the ocean floor and showing that “equatorial beams” were a natural outcome of large-scale, low-frequency equatorial wind forcing. In that study, equatorial Kelvin waves propagated downward from near the surface at an angle ω/N to the horizontal just as one would expect for internal gravity waves. In section 4 we show that, for the parameters appropriate to interannual and lower-frequency forcing, the wind-generated energy propagating downward is at such a small angle ω/N to the horizontal that it does not reach the deep-sea ocean floor before it reaches the deep-sea eastern ocean boundary. Since no baroclinic energy reaches the equatorial eastern ocean boundary deep-sea floor, we would expect that u′ = p′ = 0 there. In the next subsection we test this idea with assimilated equatorial data. Note that at interannual and lower frequencies, the flow perpendicular to the boundary is quasigeostrophic (Clarke and Shi 1991), so if p′ = 0 at a given depth at the equator, then it remains so along the deep-sea eastern ocean boundary. Consequently, we should expect p′ = 0 at the deep-sea ocean floor off California, consistent with the observed faster RW speed there and the discussion in section 2.

Fig. 6.
Fig. 6.

(top) Regression of anomalous July–November surface wind (arrows) and outgoing longwave radiation (OLR) onto an equatorial SST anomaly index (anomalous July–November SST averaged over 6°S–6°N, 180°–80°W). Wind vectors are only shown for those grid points whose u or υ correlations with the SST index exceed 0.4 in absolute value. The OLR contour interval is 10 W m−2 °C−1 of the SST index. The zero contour is darkened, and the positive contour is dashed. Values less than −20 W m−2 °C−1 are shaded. The wind regressions are based on the period 1946–85 and the OLR regression on 1974–89 (1978 missing). (bottom) As in the top panel, but for December–February [redrawn from Deser and Wallace (1990)].

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0024.1

d. Checking p = 0 at the deep-sea eastern equatorial ocean bottom boundary

We computed p′ by integrating the hydrostatic balance pz=gρ with surface boundary condition p′(0) = ρref′ where η′ is the interannual anomalous surface height displacement, ρref is a constant reference density, and g is the acceleration due to gravity. The interannual density anomaly ρ′ was computed using the nonlinear, 75-term equation of state computed by TEOS-10 thermodynamic subroutines and by removing the seasonal cycle. To calculate ρ′ we need long records of temperature and salinity within the equatorial Pacific that extend throughout the entire depth. The best available in situ observations come from the TAO/TRITON buoy network, but observations are limited to the upper 500 m. Argo float data are another possibility, but their maximum depth of observation is typically half of the average depth of the deepest parts of the equatorial Pacific, and their spatial sampling is irregular. Thus, we are left with choosing a reanalysis that best captures the complete vertical structure of temperature and salinity based on robust methods of data assimilation that incorporate the aforementioned and other observational networks. We used HYCOM’s global reanalysis (https://www.hycom.org/data/glbu0pt08/expt-19pt1). We are aware that most reanalyses and numerical models poorly resolve equatorial deep jets, and consequently the velocity fields at the equator. But this should not greatly influence the outcome of the interannual anomalies in T′ and S′ given the data assimilation methods used in HYCOM for these variables (see documentations within https://www.hycom.org/data-assimilation), and HYCOM’s robust representation of low-frequency EKW characteristics (Rydbeck et al. 2019). HYCOM reanalysis data are reported in 3-hourly increments and were averaged into monthly intervals enabling a smooth removal of the season cycle. Interannual anomalies in p′ were then filtered using an 11-point symmetric low-pass filter (Trenberth 1984) that well captures the interannual signal associated with El Niño.

To check the validity of the equatorial p′ = 0 bottom boundary condition near the deep-sea eastern equatorial boundary, we estimated equatorial p′(z, t) near the eastern equatorial Pacific boundary at two deep-sea equatorial longitudes 96° and 84°W using the above dataset. We chose longitude 96°W based on its deep-sea location and proximity to the TAO/TRITON equatorial buoy at 95°W, and longitude 84°W because it is closer to the eastern boundary and its depth is similar. An EOF analysis of the p′(z, t) data at each location showed that in both locations more than 93% of the variance was described by the first EOF mode, that is, p′(z, t) ≈ EOF1(z)φ1(t) where φ1(t) is the first principal component. Since the first EOF is such an excellent approximation to p′(z, t) at both locations, we can check the validity of the theoretical bottom boundary conditions p′ = 0 and w′ = 0 using the EOF1(z)φ1(t) approximation for p′(z, t). This implies that we can use the vertical structure function EOF1(z) to assess the vertical structure of p′(z, t). Figure 7 shows (see the blue curve in each panel) that, in both locations, EOF1(z) has a very large amplitude near the surface that decreases rapidly with depth over the top few hundred meters and is negligible at the ocean floor. This suggests that the p′ = 0 bottom boundary condition is an excellent approximation.

Fig. 7.
Fig. 7.

First vertical EOF structure functions for the interannual pressure anomalies p′(z, t) in the equatorial Pacific at 96° and 84°W from January 1996 to July 2012. Also shown are the theoretical first vertical mode structure functions for the p′ = 0 (red) and the w′ = 0 (green) bottom boundary conditions. These theoretical modes were computed using WOA13-V2 climatological data. All plots are nondimensional and the theoretical modes are a best fit to the EOF structure (see text).

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0024.1

Calculations of the first vertical mode theoretical structures for the interannual pressure are shown in the right-hand side panels of Fig. 7 for the w′ = 0 bottom boundary condition (green curve) and p′ = 0 bottom boundary condition (red curve). Neither theoretical mode structure compares well with EOF1(z) in the upper part of the water column, and this is to be expected since the interannual variability there is associated with a shallow equatorial Kelvin wave beam of energy described by many vertical modes (see section 4). However, at the bottom the w′ = 0 first mode structure (green curves) is a poor match with EOF1(z) compared to the p′ = 0 first mode structure (red curves).

The above analysis has shown that to a first approximation the interannual and lower-frequency equatorial deep-sea coastal boundary pressure signal penetrates north and south along the boundary, and in the Northern Hemisphere reaches the California Current region. At the equator the assimilated data suggests that the signal has p′ = 0 at the deep-sea floor at the eastern ocean boundary, and furthermore that the interannual pressure field at the boundary is better approximated by the first vertical mode with p′ = 0 rather than w′ = 0 as the bottom boundary condition. The dominance of the p′ = 0 first vertical mode is consistent with the dominance of this mode at the deep-sea boundary and the interannual and lower-frequency RW speeds observed off California. To complete the proposed mechanism, it remains to confirm the physics of section 3c using some equatorial wave calculations, which will show that wind-driven energy at ENSO frequencies at the equator does not reach the ocean floor before it reaches the eastern ocean boundary. We will discuss this next in section 4.

4. Low-frequency wind-forced equatorial dynamics and vertically propagating waves

a. Theoretical background

When the equatorial deep ocean is forced at the surface by the wind, the barotropic response is negligible because the barotropic response involves moving the total ocean water column, so much deeper than the mixed layer depth. The depth-dependent baroclinic energy, which dominates the ocean response, does not instantly reach the bottom, but propagates vertically. To model this simply, and so estimate whether this energy reaches the ocean bottom before it strikes an eastern or western ocean boundary, we use the linear, wind-forced, continuously stratified, equatorial ocean model of Gill and Clarke (1974) that is derived and summarized in Clarke and Liu (1993) and chapter 5 of Clarke (2008). The solution consists of a sum of a complete set of vertical modes Fn(z) that satisfy the linearized surface condition w′ = ηt and the bottom boundary condition w′ = 0 on the constant depth ocean floor z = −H. As mentioned above, the barotropic (n = 0) mode contributes negligibly to the wind-forced response, and here we concentrate on the baroclinic modes for which w′ = 0 (and consequently Fnz = 0) at the free surface z = 0 is an excellent approximation. Note that in this paper the subscript n on a variable refers to the nth baroclinic mode and the x, y, z, and t subscripts to differentiation by the x, y, z, and t variables, respectively. These variables have their usual coordinate meanings of distance eastward, distance north of the equator, distance upward from the ocean at rest and time, respectively.

The solution for the perturbation pressure p′ divided by representative density ρ*, the eastward u velocity, and the northward υ velocity have the form
(p/ρ*uυ)=n=0[pn(x,y,t)un(x,y,t)υn(x,y,t)]Fn(z),
where Fn(z) satisfies the Sturm–Liouville eigenvalue problem
(FnzN2)z+Fncn2=0,
with
Fnz=0atz=0,H.
For the low-frequency, large zonal spatial scales of interest near the equator, the un(x, y, t), υn(x, y, t), and pn(x, y, t) satisfy
untβyυn=pnx+Xn,
βyun=pny,
pntcn2+unx+υny=0
where the Coriolis parameter f = βy. In (4.4)
Xn=bnτx/ρ*,
where τx is the eastward wind stress anomaly. The forcing coefficient bn is
bn=Hmix0Fn(z)dzHmixH0Fn2(z)dz,
where Hmix is the mixed layer depth, and it has been assumed that the turbulent stress divergence due to the wind stress forcing is independent of z and equal to τx/Hmix in the mixed layer and zero beneath the mixed layer.
For each vertical mode n the solution of (4.4)(4.6) can be expressed in terms of a complete infinite set of Hermite functions ψm(y). Physically, these modes represent long equatorial RW modes and one EKW mode. Of these waves the EKW is the only equatorial wave with eastward group velocity, and so this is the only wave that can carry energy to the eastern equatorial boundary of the Pacific Ocean, and so determine whether p′ = 0 at the ocean floor. The EKW solution for each vertical mode n is of the form
unKel=pnKelcn=π0.25qn(x,t)Fn(z)exp(βy22cn),
the y dependence being proportional to zeroth-order Hermite function ψ0[y(β/cn)1/2], and qn having the dimensions of a velocity. The function qn(x, t) satisfies a wind-forced EKW problem that is the dimensional form of Eq. (A9) of Clarke and Liu (1993) or (5.11) of Clarke (2008):
qnt+cnqnx=21π0.25+bnτxρ*1exp(0.5ξ2)dξ,
where ξ=yβ/cn.

b. Application to the equatorial Pacific

From (4.10) the EKW can be calculated without separating into its various frequencies, but in order to illustrate the vertically propagating energy as “rays,” we write the interannual equatorial wind forcing of interest in terms of its frequency components ω. Since also there is only weak zonal forcing in the eastern equatorial Pacific (see Fig. 6), we will write, for frequency ω,
τx={Re{τoxeiωt},fora<x<0,0,otherwise,
where τox is a constant, x = 0 corresponds to 150°W and x = −a is 150°E. Likewise, the eastern and western equatorial boundaries are defined as x = b (80°W) and x = −d (130°E), respectively. This is an idealized representation of the wind forcing observed in the growth stage of El Niño events (see Fig. 6, top). For all of the baroclinic vertical modes n, the EKW meridional scale (2cn/β)0.5 is smaller than 5° of latitude and, for most modes n, it is much smaller. In other words, for most vertical modes the wind forcing is essentially constant meridionally. To keep our illustration as simple as possible, we have therefore assumed that τox is independent of y and constant zonally in the region of wind stress forcing.
Since within the forcing region τox is independent of y, the integral in (4.10) is known analytically and the function qn satisfies
qnt+cnqnx=20.5bnτ0xρ*1eiωt.
In (4.12) and following we will use complex variables and take the real part at the end to get the solution. If we only consider the eastward-propagating EKWs directly forced by the wind stress, qn = 0 at x = −a, the western end of the forcing region. Hence the solution of (4.12) in the forcing region is
qn=20.5ibnτ0xρ*1ω1[eiωteiωtiω(x+a)cn1].
The solution east of the region of forcing is a freely propagating EKW that matches this solution at x = 0 and so is
qn=20.5ibnτ0xρ*1ω1[eiωteiωxcn1eiωtiω(x+a)cn1].
Because the solution (4.13) and (4.14) uses qn = 0 at x = −a, it ignores the EKW energy resulting from equatorial RWs that are generated in the forcing region, propagate westward, and reflect from the western equatorial ocean boundary. The EKWs resulting from this RW reflection at the western ocean boundary propagate eastward as unforced Kelvin waves right across the equatorial Pacific to the eastern ocean boundary. In theory these waves will also contribute to the signal at the eastern boundary. However, McCreary’s (1984) analysis showed that, even for a western boundary that is a meridional wall, the reflected equatorial wave energy is much weaker than that for the EKW directly forced by the wind. In addition, because of the dissipation of energy at the irregular western boundary, especially for the higher-order vertical modes, the main eastward-propagating EKW signal east of the region of wind forcing x = 0 is the signal directly generated by the wind stress shown in the upper panel of Fig. 6.

c. Summation of vertical modes to obtain EKW rays

To illustrate how the sum of the baroclinic modes can produce downward-propagating “rays,” the eigenfunctions Fn(z) will be approximated analytically by the WKB eigenfunctions for which the buoyancy frequency N(z) varies more slowly than the phase of the vertical modes (see, e.g., Clarke and Van Gorder 1986). Strictly speaking, this is not valid for the lowest vertical modes, but numerically it is accurate enough that the analytical solutions we obtain illustrate the essential physics. For the w′ = 0 bottom boundary condition, Fn(z) can be written
Fn(z)=[N(z)N(0)]0.5cos[S(z)cn],
where
S(z)=z0N(z*)dz*
and
cn=c1/n=N¯H/nπ,
where N¯ is the average value of N(z) over the water depth H and N(0) is the average weak vertical stratification at the surface. Note that the only difference for Fn(z) for the p′ = 0 bottom boundary condition modes is that Fn(−H) = 0 and hence that
cn=N¯H(n0.5)π
Equations (4.18) and (4.17) show that c1 for the p′ = 0 bottom boundary condition is double the corresponding c1 for the w′ = 0 bottom boundary condition case. This is qualitatively consistent with the faster RW speeds for p′ = 0 for the first vertical mode off California in Fig. 1, but it is not numerically accurate because (4.17) and (4.18) are only valid for large n. From (4.15) and (4.17), cos(S/cn) = cos[nS(z)/c1] = [exp(inS/c1) + exp(−inS/c1)]/2 and hence that
Fn(z)=[N(z)/N(0)]1/2[exp(inS/c1)+exp(inS/c1)]/2.
The terms being summed in (4.9) have x dependence proportional to exp(−inωx/c1) = exp(−iωx/c1)n, a y dependence proportional to {exp[βy2/(2c1)]}n, and from (4.19) a z dependence proportional to exp(iS/c1)n + exp(−iS/c1)n. The mixed layer depth is much smaller than the water depth, and (see appendix) in the limit of an infinitesimally thin surface mixed layer the forcing coefficient bn=2N(0)/(N¯H). Based on (4.14), the above exponential results and the results for bn, we have, for x ≥ 0,
qn=Aeiωt([exp(iωxc1)]n+{exp[iω(x+a)c1]}n),
where
A=20.5iτ0xN(0)(N¯Hρ*ω)1.
Substitution of (4.19), (4.20), and {exp[βy2/(2c1)]}n=Rn into (4.9) implies that uKel can be written
uKel=0.5π0.25Aeiωt[N(z)N(0)]0.5[n=1(RM)n+(RQ)n+(Rμ)n+(Rζ)n],
where M=e(iωx+iS)/c1, Q=e(iωxiS)/c1, μ=Meiωa/c1, and ζ=Qeiωa/c1. Summing the geometric series in (4.22) gives
uKel=0.5π0.25A[N(z)N(0)]0.5eiωt(RM1RM+RQ1RQ+Rμ1Rμ+Rζ1Rζ).
Notice that uKel has singularities where RM, RQ, , or = 1. These singularities can only occur at the equator y = 0 where R = 1 and where M, Q, μ or ζ = 1. Based on the expressions for M, Q, μ, and ζ the singularities occur when −ωx ± S = 2πjc1 or −ω(x + a) ± S = 2πjc1, for any integer j, that is,
x=±Sω2πjc1ωa,x=±Sω2πjc1ω.
It follows from (4.16) that S is a nonnegative decreasing function of z, being N¯H at z = −H and 0 at z = 0. From (4.17) c1=N¯H/π, so at z = −H the location of the singularity with the smallest positive x in (4.24) is found from
x=±c1πω2πjc1ωa.
Note that c1π/ω can also be written as 0.5c1T where T is the period of the forcing. For c1 = 2.5 m s−1 and T = 2 years, the smallest reasonable values for interannual forcing, c1π/ω corresponds to 353° of longitude. Since the region of forcing stretches from 150°E (x = −a) to 150°W (x = 0), a = 60° of longitude, and consequently from (4.25) the smallest possible positive value of x for which the singularity reaches the bottom occurs for j = 0 when x = +c1π/ωa = 353 − 60 = 293° of longitude. This is much further east than the eastern Pacific Ocean boundary [about 70° of longitude east of 150°W (x = 0)]. In other words, for idealized, but realistic parameters, the singularity does not reach the ocean floor before it reaches the eastern boundary.

Note from (4.24) for shallower depths and j = 0, x = S/ωa is small enough for singularities to occur before the eastern equatorial Pacific boundary. From (4.23) this corresponds to a vertical phase propagation proportional to exp(iωt + iS/c1), that is, from the definition of S, upward phase propagation at the WKB slowly varying speed c1ω/N(z). Since the phase propagation is also c1 eastward, the phase velocity is upward and eastward with slope +ω/N(z) to the horizontal.

The above analysis focused on uKel, but the same singularities and corresponding results also occur for pKel for which, for mode n [see (4.9)] pnKel=cnunKel=c1unKel/n. The infinite sums corresponding to (4.21) for pKel are therefore of the form
n=1[(RM)n+(RQ)n+(Rμ)n+(Rζ)n]c1n=c1[ln(1RM)+ln(1RQ)+ln(1Rμ)+ln(1Rζ)],
which has logarithmic singularities at the same locations as uKel.
Physically, the above singularities correspond to freely propagating internal Kelvin waves that must have downward group velocity away from the surface. Internal Kelvin waves are really inertia gravity waves trapped by rotation to the equator and propagate at an angle −ω/N to the horizontal. Internal gravity waves have the vertical component of their group velocity equal and opposite to their phase velocity, so for energy propagating downward the corresponding phase velocity is upward at an angle ω/N to the horizontal as we found earlier. The relevant equation for the singularity has the positive choice for the sign in (4.24), so differentiation of (4.24) with respect to z enables us to determine the slope of the singularity and energy propagation to the horizontal at the expected angle
dzdx=ωN(z).
Note that we allow the buoyancy frequency N to vary slowly along the equator, so, strictly speaking, N is also a slowly varying function of x as well as z.

Figure 8 shows the ray paths of EKWs emitted at x = 0 (150°W). We assumed that at x = 0 the EKWs begin their journey at 50-m depth, the approximate climatological depth of the surface (mixed layer Ando and McPhaden 1997; de Boyer Montégut et al. 2004; Zhang and Clarke 2015). The plot for ω = 2π/2 years (solid white line) shows that even at this “high” interannual frequency the EKWs beginning at x = 0 only reach 280-m depth. For lower interannual frequencies the water depths are even shallower, far above the ocean floor. For example, the dotted white line in Fig. 8, right above the ω = 2π/2 years ray path, represents the ray path for ω = 2π/3 years, which has incidence at about 190-m depth. Therefore, it is not surprising that the p′ = 0 bottom boundary condition for the baroclinic modes is valid at interannual frequencies. This is consistent with other results that show that wind-forced vertically propagating baroclinic EKWs approximately follow the angle ω/N(z) from the surface (see, e.g., McCreary 1984; Dewitte and Reverdin 2000; Kutsuwada and McPhaden 2002; Rydbeck et al. 2019) despite experiencing variable wind forcing and complex ocean currents before reaching the eastern ocean boundary. It is also consistent with the results obtained from the assimilated data in section 3d.

Fig. 8.
Fig. 8.

Integrated EKW WKB ray paths beginning at the climatological mixed layer depth of 50 m at 150°W (de Boyer Montégut et al. 2004; Zhang and Clarke 2015). Shaded contours in the background represent the climatological buoyancy frequency as a function of depth and longitude averaged across the latitude band 2°S–2°N. The longitudinal domain extends from the eastern edge of interannual wind forcing (150°W) in the equatorial Pacific associated with El Niño events (see Fig. 6, top panel) to its eastern boundary (80°W). WKB ray paths reveal the depth of penetration for EKWs, whether forced in the interior or emitted from the western boundary, with dependence on angular frequency: ω = 2π/2 years (solid white line) and ω = 2π/3 years (dotted white line).

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0024.1

5. Finite-thickness mixed layer and the dominance of the first vertical mode

When the surface mixed layer is finite, the singular “rays” at the equator are no longer infinitesimally thick but have a vertical scale comparable to the mixed layer depth, and are more appropriately called equatorial “beams.” This has been shown numerically by McCreary (1984). Mathematically, when the mixed layer thickness is finite bn is no longer independent of n, but decreases rapidly with increasing n, so that the ocean response is dominated by the lower-order modes. For example, in the formulation by Zhang and Clarke (2017), bn decreases like n−2. This, and the stronger shears and dissipation associated with the higher vertical modes, is the likely reason for the dominance of the first vertical mode seen observationally (see Fig. 7) in the interannual pressure field.

6. Vertical wave propagation, bottom roughness, and negligible bottom pressure

As mentioned in the third paragraph of the introduction, previous analysis has suggested that the ocean floor is dynamically rough, and such bottom topography can explain why p′ = 0 on the ocean floor and the consequent increase in wave Rossby wave speed. We have suggested that this is not the physical reason for the faster Rossby wave speed off California, because the interannual wave variability has not yet reached the ocean floor there and therefore has not interacted with the rough bottom topography. What we have not done is to check how far from the coast our mechanism is valid, and to what extent the vertical propagation mechanism for bottom p′ = 0 applies elsewhere.

a. How far does the Rossby wave propagate westward and vertically before reaching the ocean floor?

Our analysis has shown that the interannual surface pressure field near the eastern boundary is dominated by the first vertical mode with zero bottom pressure. Previous numerical and analytical calculations (Dewitte and Reverdin 2000; Ramos et al. 2008; Vergara et al. 2017) suggest that the interannual variability propagates westward from the coast as vertically propagating Rossby waves in the form of wave beams with slope given for the nth vertical mode by
dz/dx=2ωf2/(βcnN).
This can also be written as
dx/dz=βcnN/(2ωf2),
and integrated from x(−H) where the wave ray first touches the bottom to the deep-sea “coast” at x = 0 to give
x(H)=0.5βcnω1f2H0N(z)dz.
We evaluated (6.3) for interannual variability off California with the appropriate values f=0.79×104s1, β=1.9×1011m1s1, H0N(z)dz=8.88ms1, and the representative interannual frequency ω = 2π/3.3 years. We used the dominant first vertical mode phase speed c1 = 3.68 m s−1 corresponding to the p′ = 0 bottom boundary condition for the first vertical mode with u¯=0 and Rossby wave speed βc12f2=4.12cms1 (see Table 1, third column). This p′ = 0 bottom boundary condition is valid before the vertically propagating nonzero p′ Rossby wave energy reaches the bottom and is valid at x(−H) because of the rough bottom. With these parameter values in (6.3) we determine that the zonal location of where the Rossby wave first strikes the bottom is (−H) = −824 km. This is approximately the same zonal distance as the zonal width of the California Current region analyzed (see the vertical axes in Fig. 2). Thus p′ = 0 on the ocean floor in the California Current region is mostly due to Rossby wave energy not reaching the bottom.

b. Application to other latitudes and frequencies

Equation (6.3) shows that the extent of influence of the vertically propagating Rossby waves increases like β f−2, so although x(−H) is only about 824 km at 33°N, at 5° latitude x(−H) is wider than the Pacific Ocean basin. This is also true within the 5°S to 5°N equatorial waveguide, for there vertically propagating equatorial Kelvin waves and first meridional mode equatorial Rossby waves dominate the ocean response and cross the Pacific without striking the bottom. Evidence for this is provided by Dewitte and Reverdin (2000). Note that the Indian and Atlantic Ocean basins are much narrower than the Pacific, and so x(−H) does not have to be as large to guarantee that the interannual Rossby waves do not reach the ocean floor before being stopped by a boundary. In those basins most of the interannual Rossby wave variability will not reach the ocean floor in the latitude band from 15°S to 15°N. Note that Eq. (6.3) shows that x(−H) is proportional to ω−1, so the range of latitudes for which decadal variability does not reach the ocean floor is about 3 times that of interannual variability.

7. Summary and concluding remarks

Our goal in this paper has been to understand physically why longwave RWs propagating westward from the California coast propagate at a much faster speed than we would expect. We also wanted to understand why this propagation speed agrees with the first vertical mode RW speed with the w′ = 0 bottom boundary condition replaced by the p′ = 0 bottom boundary condition (Tailleux and McWilliams 2001).

Past work, consistent with our updated analysis in section 3, has shown that long RWs propagating westward from the California coast at interannual and lower frequencies result from incident EKWs, the only way such energy can reach the eastern ocean boundary from the interior ocean. The baroclinic EKW energy propagates at such a shallow angle ω/N(z) to the vertical that it is unlikely to reach the ocean bottom before it encounters the deep-sea eastern boundary of the equatorial Pacific. Consequently, p′ = 0 is an appropriate bottom boundary condition at the equator at the deep-sea eastern Pacific boundary, a result consistent with our analysis of the assimilation data in section 3d and the approximation of these data by the first vertical mode with the w′ = 0 bottom boundary condition replaced by p′ = 0. Theory suggests (see Clarke and Shi 1991) that the motion at these low frequencies is quasigeostrophic perpendicular to the boundary, and so argues for a p′ = 0 bottom boundary condition at the deep-sea boundary off California. Our updated detailed analysis in section 2 confirms that such a deep-sea boundary condition and associated first vertical mode gives offshore RWs that propagate westward at the observed speed.

Data availability statement.

The data we used are available from the following sources: SSH—Zlotnicki et al. (2019); hydrographic data—Locarnini et al. (2013) and Zweng et al. (2013); monthly sea level—Permanent Service for Mean Sea Level (PSMSL; http://www.psmsl.org/data/obtaining/); monthly Niño-3.4 index—https://stateoftheocean.osmc.noaa.gov/sur/pac/nino34.php; HYCOM assimilation data—https://www.hycom.org/data/glbu0pt08/expt-19pt1.

APPENDIX

WKB Expression for Wind Forcing Coefficient bn and Its Hmix → 0 Limit

The wind forcing coefficient bn in (4.8) can be evaluated using the WKB approximations for Fn(z) given in (4.15)(4.17). As Hmix approaches zero, the numerator divided by Hmix is given by Hmix1HmixFn(0)=1 since Fn(0) = 1 by (4.15). The integral in the denominator of (4.8), using (A2) and (A7) from Zhang and Clarke (2017) is
H0Fn2(z)dz=0.5N¯HN(0).
Therefore, the wind forcing coefficient
bn=2N(0)HN¯

REFERENCES

  • Ando, K., and M. J. McPhaden, 1997: Variability of surface later hydrography in the tropical Pacific Ocean. J. Geophys. Res., 102, 23 06323 078, https://doi.org/10.1029/97JC01443.

    • Search Google Scholar
    • Export Citation
  • Bobrovich, A. V., and G. M. Reznik, 1999: Planetary waves in a stratified ocean of variable depth. Part 2. Continuously stratified ocean. J. Fluid Mech., 388, 147169, https://doi.org/10.1017/S0022112099004863.

    • Search Google Scholar
    • Export Citation
  • Bretschneider, D. F., and D. R. McLain, 1976: Anomalies of monthly mean sea level along the west coasts of North and South America. Mar. Fish. Climate, 3, 95102.

    • Search Google Scholar
    • Export Citation
  • Brink, K. H., and J. Pedlosky, 2018: Rossby waves with continuous stratification and bottom friction. J. Phys. Oceanogr., 48, 22092219, https://doi.org/10.1175/JPO-D-18-0070.1.

    • Search Google Scholar
    • Export Citation
  • Brink, K. H., and J. Pedlosky, 2020: The structure of baroclinic modes in the presence of baroclinic mean flow. J. Phys. Oceanogr., 50, 239253, https://doi.org/10.1175/JPO-D-19-0123.1.

    • Search Google Scholar
    • Export Citation
  • Cane, M. A., and E. S. Sarachik, 1977: Forced baroclinic ocean motions. II. The equatorial bounded case. J. Mar. Res., 35, 395432.

  • Cane, M. A., and D. W. Moore, 1981: A note on low-frequency equatorial basin modes. J. Phys. Oceanogr., 11, 15781584, https://doi.org/10.1175/1520-0485(1981)011<1578:ANOLFE>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Challenor, P. G., P. Cipollini, D. Cromwell, K. L. Hill, G. D. Quartly, and I. S. Robinson, 2004: Characteristics of mid-latitude Rossby wave propagation from multiple satellite datasets. Int. J. Remote Sens., 25, 12971302, https://doi.org/10.1080/01431160310001592201.

    • Search Google Scholar
    • Export Citation
  • Chelton, D. B., 1980: Low frequency sea level variability along the west coast of North America. Ph.D. dissertation, Scripps Institution of Oceanography, University of California, 212 pp.

  • Chelton, D. B., and R. E. Davis, 1982: Monthly mean sea-level variability along the west coast of North America. J. Phys. Oceanogr., 12, 757784, https://doi.org/10.1175/1520-0485(1982)012<0757:MMSLVA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Chelton, D. B., P. A. Bernal, and J. A. McGowan, 1982: Large-scale interannual physical and biological interaction in the California Current. J. Mar. Res., 40, 10951125.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., 1983: The reflection of equatorial waves from oceanic boundaries. J. Phys. Oceanogr., 13, 11931207, https://doi.org/10.1175/1520-0485(1983)013<1193:TROEWF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., 1992: Low-frequency reflection from a nonmeridional eastern ocean boundary and the use of coastal sea level to monitor eastern Pacific equatorial Kelvin waves. J. Phys. Oceanogr., 22, 163183, https://doi.org/10.1175/1520-0485(1992)022<0163:LFRFAN>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., 2008: An Introduction to the Dynamics of El Niño and the Southern Oscillation. Academic Press, 324 pp.

  • Clarke, A. J., and R. O. R. Y. Thompson, 1984: Large-scale wind driven ocean response in the Australian Coastal Experiment region. J. Phys. Oceanogr., 14, 338352, https://doi.org/10.1175/1520-0485(1984)014<0338:LSWDOR>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., and S. Van Gorder, 1986: A method for estimating wind-driven frictional, time-dependent, stratified shelf and slope water flow. J. Phys. Oceanogr., 16, 10131028, https://doi.org/10.1175/1520-0485(1986)016<1013:AMFEWD>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., and C. Shi, 1991: Critical frequencies at ocean boundaries. J. Geophys. Res., 96, 10 73110 738, https://doi.org/10.1029/91JC00933.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., and X. Liu, 1993: Observations and dynamics of semiannual and annual sea levels near the eastern equatorial Indian Ocean boundary. J. Phys. Oceanogr., 23, 386399, https://doi.org/10.1175/1520-0485(1993)023<0386:OADOSA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., and A. Lebedev, 1999: Remotely driven decadal and longer changes in the coastal Pacific waters of the Americas. J. Phys. Oceanogr., 29, 828835, https://doi.org/10.1175/1520-0485(1999)029<0828:RDDALC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., and M. Dottori, 2008: Planetary wave propagation off California and its effect on zooplankton. J. Phys. Oceanogr., 38, 702714, https://doi.org/10.1175/2007JPO3691.1.

    • Search Google Scholar
    • Export Citation
  • de Boyer Montégut, C., G. Madec, A. S. Fischer, A. Lazar, and D. Iudicone, 2004: Mixed layer depth over the global ocean: An examination of profile data and a profile-based climatology. J. Geophys. Res., 109, C12003, https://doi.org/10.1029/2004JC002378.

    • Search Google Scholar
    • Export Citation
  • de La Lama, M. S., L. H. LaCasce, and H. Fuhr, 2016: The vertical structure of ocean eddies. Dyn. Stat. Climate Syst., 1, dzw001, https://doi.org/10.1093/climsys/dzw001.

    • Search Google Scholar
    • Export Citation
  • Deser, C., and J. M. Wallace, 1990: Large-scale atmospheric circulation features of warm and cold episodes in the tropical Pacific. J. Climate, 3, 12541281, https://doi.org/10.1175/1520-0442(1990)003<1254:LSACFO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Dewitte, B., and G. Reverdin, 2000: Vertically propagating annual and interannual variability in an OGCM simulation of the tropical Pacific Ocean in 1985–94. J. Phys. Oceanogr., 30, 15621581, https://doi.org/10.1175/1520-0485(2000)030<1562:VPAAIV>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Dottori, M., and A. J. Clarke, 2009: Rossby waves and the interannual and interdecadal variability of temperature and salinity off California. J. Phys. Oceanogr., 39, 25432561, https://doi.org/10.1175/2009JPO3898.1.

    • Search Google Scholar
    • Export Citation
  • Ebisuzaki, W., 1997: A method to estimate the statistical significance of a correlation when the data are serially correlated. J. Climate, 10, 21472153, https://doi.org/10.1175/1520-0442(1997)010<2147:AMTETS>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Enfield, D. B., and J. S. Allen, 1980: On the structure and dynamics of monthly mean sea level anomalies along the Pacific coast of North and South America. J. Phys. Oceanogr., 10, 557578, https://doi.org/10.1175/1520-0485(1980)010<0557:OTSADO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Frischknecht, M., M. Münnich, and N. Gruber, 2015: Remote versus local influence of ENSO on the California Current System. J. Geophys. Res. Oceans, 120, 13531374, https://doi.org/10.1002/2014JC010531.

    • Search Google Scholar
    • Export Citation
  • Fu, L.-L., and B. Qiu, 2002: Low-frequency variability of the North Pacific Ocean: The role of boundary- and wind-driven baroclinic Rossby waves. J. Geophys. Res., 107, 3220, https://doi.org/10.1029/2001JC001131.

    • Search Google Scholar
    • Export Citation
  • Gill, A. E., and A. J. Clarke, 1974: Wind-induced upwelling, coastal currents and sea-level changes. Deep-Sea Res. Oceanogr. Abtsr., 21, 325345, https://doi.org/10.1016/0011-7471(74)90038-2.

    • Search Google Scholar
    • Export Citation
  • Giunipero, E., and A. J. Clarke, 2013: Estimation of the effect of eddies on coastal El Niño flows using along-track satellite altimeter data. J. Phys. Oceanogr., 43, 12091224, https://doi.org/10.1175/JPO-D-12-0109.1.

    • Search Google Scholar
    • Export Citation
  • Grimshaw, R., and J. S. Allen, 1988: Low-frequency baroclinic waves off coastal boundaries. J. Phys. Oceanogr., 18, 11241143, https://doi.org/10.1175/1520-0485(1988)018<1124:LFBWOC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Herrera Cervantes, H., and A. Parés-Sierra, 1994: Propagación de variaciones de baja frecuencia en la temperatura superficial del Pacífico nor-oriental. Geofis. Int., 33, 469486, https://doi.org/10.22201/igeof.00167169p.1994.33.3.1184.

    • Search Google Scholar
    • Export Citation
  • Kessler, W. S., 1990: Observations of long Rossby waves in the northern tropical Pacific. J. Geophys. Res., 95, 51835217, https://doi.org/10.1029/JC095iC04p05183.

    • Search Google Scholar
    • Export Citation
  • Killworth, P. D., and J. R. Blundell, 1999: The effects of bottom topography on the speed of long extratropical planetary waves. J. Phys. Oceanogr., 29, 26892710, https://doi.org/10.1175/1520-0485(1999)029<2689:TEOBTO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Killworth, P. D., and J. R. Blundell, 2003: Long extratropical planetary wave propagation in the presence of slowly varying mean flow and bottom topography. Part I: The local problem. J. Phys. Oceanogr., 33, 784801, https://doi.org/10.1175/1520-0485(2003)33<784:LEPWPI>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Killworth, P. D., D. B. Chelton, and R. A. de Szoeke, 1997: The speed of observed and theoretical long extratropical planetary waves. J. Phys. Oceanogr., 27, 19461966, https://doi.org/10.1175/1520-0485(1997)027<1946:TSOOAT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Kutsuwada, K., and M. McPhaden, 2002: Intraseasonal variations in the upper equatorial Pacific Ocean prior to and during the 1997–98 El Niño. J. Phys. Oceanogr., 32, 11331149, https://doi.org/10.1175/1520-0485(2002)032<1133:IVITUE>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • LaCasce, J. H., 2017: The prevalence of oceanic surface modes. Geophys. Res. Lett., 44, 11 09711 105, https://doi.org/10.1002/2017GL075430.

    • Search Google Scholar
    • Export Citation
  • LaCasce, J. H., and S. Groeskamp, 2020: Baroclinic modes over rough bathymetry and the surface deformation radius. J. Phys. Oceanogr., 50, 28352847, https://doi.org/10.1175/JPO-D-20-0055.1.

    • Search Google Scholar
    • Export Citation
  • Li, J., and A. J. Clarke, 2004: Coastline direction, interannual flow and the strong El Niño currents along Australia’s nearly zonal southern coast. J. Phys. Oceanogr., 34, 23732381, https://doi.org/10.1175/JPO2645.1.

    • Search Google Scholar
    • Export Citation
  • Locarnini, R. A., and Coauthors, 2013: Temperature. Vol. 1, World Ocean Atlas 2013, NOAA Atlas NESDIS 73, 40 pp., http://data.nodc.noaa.gov/woa/WOA13/DOC/woa13_vol1.pdf.

  • McCreary, J. P., 1977: Eastern ocean response to changing wind systems. Ph.D. dissertation, Scripps Institution of Oceanography, University of California, 156 pp.

  • McCreary, J. P., 1984: Equatorial beams. J. Mar. Res., 42, 395430, https://doi.org/10.1357/002224084788502792.

  • Ramos, M., B. Dewitte, O. Pizzaro, and G. Garric, 2008: Vertical propagation of extratropical Rossby waves during the 1997–1998 El Niño off the west coast of South America in a medium-resolution OGCM simulation. J. Geophys. Res., 113, C08041, https://doi.org/10.1029/2007JC004681.

    • Search Google Scholar
    • Export Citation
  • Ramp, S. R., J. L. McClean, C. A. Collins, A. J. Semtner, and K. A. S. Hays, 1997: Observations and modeling of the 1991–1992 El Niño signal off central California. J. Geophys. Res., 102, 55535582, https://doi.org/10.1029/96JC03050.

    • Search Google Scholar
    • Export Citation
  • Rydbeck, A. V., T. G. Jensen, and M. Flatau, 2019: Characterization of intraseasonal kelvin waves in the equatorial Pacific Ocean. J. Geophys. Res. Oceans, 124, 20282053, https://doi.org/10.1029/2018JC014838.

    • Search Google Scholar
    • Export Citation
  • Samelson, R. M., 1992: Surface-intensified Rossby waves over rough topography. J. Mar. Res., 50, 367384, https://doi.org/10.1357/002224092784797593.

    • Search Google Scholar
    • Export Citation
  • Schopf, P. S., D. L. T. Anderson, and R. Smith, 1981: Beta dispersion of low-frequency Rossby waves. Dyn. Atmos. Oceans, 5, 187214, https://doi.org/10.1016/0377-0265(81)90011-7.

    • Search Google Scholar
    • Export Citation
  • Strub, P. T., and C. James, 2002: The 1997–1998 oceanic El Niño signal along the southeast and northeast Pacific boundaries—An altimetric view. Prog. Oceanogr., 54, 439458, https://doi.org/10.1016/S0079-6611(02)00063-0.

    • Search Google Scholar
    • Export Citation
  • Tailleux, R., and J. C. McWilliams, 2001: The effect of bottom pressure decoupling on the speed of extratropical, baroclinic Rossby waves. J. Phys. Oceanogr., 31, 14611476, https://doi.org/10.1175/1520-0485(2001)031<1461:TEOBPD>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Trenberth, K. E., 1984: Signal versus noise in the Southern Oscillation. Mon. Wea. Rev., 112, 326332, https://doi.org/10.1175/1520-0493(1984)112<0326:SVNITS>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Vega, A., Y. duPenhoat, B. Dewitte, and O. Pizzaro, 2003: Equatorial forcing of interannual Rossby waves in the eastern South Pacific. Geophys. Res. Lett., 30, 1197, https://doi.org/10.1029/2002GL015886.

    • Search Google Scholar
    • Export Citation
  • Vergara, O., B. Dewitte, M. Ramos, and O. Pizzaro, 2017: Vertical energy flux at ENSO time scales in the subthermoclinie of the southeastern Pacific. J. Geophys. Res. Oceans, 122, 60116038, https://doi.org/10.1002/2016JC012614.

    • Search Google Scholar
    • Export Citation
  • Wortham, C., and C. Wunsch, 2014: A multidimensional spectral description of ocean variability. J. Phys. Oceanogr., 44, 944966, https://doi.org/10.1175/JPO-D-13-0113.1.

    • Search Google Scholar
    • Export Citation
  • Wunsch, C., 1997: The vertical partition of oceanic horizontal kinetic energy. J. Phys. Oceanogr., 27, 17701794, https://doi.org/10.1175/1520-0485(1997)027<1770:TVPOOH>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Zhang, X., and A. J. Clarke, 2015: Observations of interannual equatorial freshwater jets in the western Pacific. J. Phys. Oceanogr., 45, 28482865, https://doi.org/10.1175/JPO-D-14-0245.1.

    • Search Google Scholar
    • Export Citation
  • Zhang, X., and A. J. Clarke, 2017: On the dynamical relationship between equatorial Pacific surface currents, zonally averaged equatorial sea level, and El Niño predictions. J. Phys. Oceanogr., 47, 323337, https://doi.org/10.1175/JPO-D-16-0193.1.

    • Search Google Scholar
    • Export Citation
  • Zlotnicki, V., Z. Qu, and J. Willis, 2019: MEaSUREs gridded sea surface height anomalies version 1812. PO.DAAC, accessed 22 May 2019, https://doi.org/10.5067/SLREF-CDRV2.

  • Zweng, M. M., and Coauthors, 2013: Salinity. Vol. 2, World Ocean Atlas 2013, NOAA Atlas NESDIS 74, 39 pp., http://data.nodc.noaa.gov/woa/WOA13/DOC/woa13_vol2.pdf.

Save
  • Ando, K., and M. J. McPhaden, 1997: Variability of surface later hydrography in the tropical Pacific Ocean. J. Geophys. Res., 102, 23 06323 078, https://doi.org/10.1029/97JC01443.

    • Search Google Scholar
    • Export Citation
  • Bobrovich, A. V., and G. M. Reznik, 1999: Planetary waves in a stratified ocean of variable depth. Part 2. Continuously stratified ocean. J. Fluid Mech., 388, 147169, https://doi.org/10.1017/S0022112099004863.

    • Search Google Scholar
    • Export Citation
  • Bretschneider, D. F., and D. R. McLain, 1976: Anomalies of monthly mean sea level along the west coasts of North and South America. Mar. Fish. Climate, 3, 95102.

    • Search Google Scholar
    • Export Citation
  • Brink, K. H., and J. Pedlosky, 2018: Rossby waves with continuous stratification and bottom friction. J. Phys. Oceanogr., 48, 22092219, https://doi.org/10.1175/JPO-D-18-0070.1.

    • Search Google Scholar
    • Export Citation
  • Brink, K. H., and J. Pedlosky, 2020: The structure of baroclinic modes in the presence of baroclinic mean flow. J. Phys. Oceanogr., 50, 239253, https://doi.org/10.1175/JPO-D-19-0123.1.

    • Search Google Scholar
    • Export Citation
  • Cane, M. A., and E. S. Sarachik, 1977: Forced baroclinic ocean motions. II. The equatorial bounded case. J. Mar. Res., 35, 395432.

  • Cane, M. A., and D. W. Moore, 1981: A note on low-frequency equatorial basin modes. J. Phys. Oceanogr., 11, 15781584, https://doi.org/10.1175/1520-0485(1981)011<1578:ANOLFE>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Challenor, P. G., P. Cipollini, D. Cromwell, K. L. Hill, G. D. Quartly, and I. S. Robinson, 2004: Characteristics of mid-latitude Rossby wave propagation from multiple satellite datasets. Int. J. Remote Sens., 25, 12971302, https://doi.org/10.1080/01431160310001592201.

    • Search Google Scholar
    • Export Citation
  • Chelton, D. B., 1980: Low frequency sea level variability along the west coast of North America. Ph.D. dissertation, Scripps Institution of Oceanography, University of California, 212 pp.

  • Chelton, D. B., and R. E. Davis, 1982: Monthly mean sea-level variability along the west coast of North America. J. Phys. Oceanogr., 12, 757784, https://doi.org/10.1175/1520-0485(1982)012<0757:MMSLVA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Chelton, D. B., P. A. Bernal, and J. A. McGowan, 1982: Large-scale interannual physical and biological interaction in the California Current. J. Mar. Res., 40, 10951125.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., 1983: The reflection of equatorial waves from oceanic boundaries. J. Phys. Oceanogr., 13, 11931207, https://doi.org/10.1175/1520-0485(1983)013<1193:TROEWF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., 1992: Low-frequency reflection from a nonmeridional eastern ocean boundary and the use of coastal sea level to monitor eastern Pacific equatorial Kelvin waves. J. Phys. Oceanogr., 22, 163183, https://doi.org/10.1175/1520-0485(1992)022<0163:LFRFAN>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., 2008: An Introduction to the Dynamics of El Niño and the Southern Oscillation. Academic Press, 324 pp.

  • Clarke, A. J., and R. O. R. Y. Thompson, 1984: Large-scale wind driven ocean response in the Australian Coastal Experiment region. J. Phys. Oceanogr., 14, 338352, https://doi.org/10.1175/1520-0485(1984)014<0338:LSWDOR>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., and S. Van Gorder, 1986: A method for estimating wind-driven frictional, time-dependent, stratified shelf and slope water flow. J. Phys. Oceanogr., 16, 10131028, https://doi.org/10.1175/1520-0485(1986)016<1013:AMFEWD>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., and C. Shi, 1991: Critical frequencies at ocean boundaries. J. Geophys. Res., 96, 10 73110 738, https://doi.org/10.1029/91JC00933.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., and X. Liu, 1993: Observations and dynamics of semiannual and annual sea levels near the eastern equatorial Indian Ocean boundary. J. Phys. Oceanogr., 23, 386399, https://doi.org/10.1175/1520-0485(1993)023<0386:OADOSA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., and A. Lebedev, 1999: Remotely driven decadal and longer changes in the coastal Pacific waters of the Americas. J. Phys. Oceanogr., 29, 828835, https://doi.org/10.1175/1520-0485(1999)029<0828:RDDALC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Clarke, A. J., and M. Dottori, 2008: Planetary wave propagation off California and its effect on zooplankton. J. Phys. Oceanogr., 38, 702714, https://doi.org/10.1175/2007JPO3691.1.

    • Search Google Scholar
    • Export Citation
  • de Boyer Montégut, C., G. Madec, A. S. Fischer, A. Lazar, and D. Iudicone, 2004: Mixed layer depth over the global ocean: An examination of profile data and a profile-based climatology. J. Geophys. Res., 109, C12003, https://doi.org/10.1029/2004JC002378.

    • Search Google Scholar
    • Export Citation
  • de La Lama, M. S., L. H. LaCasce, and H. Fuhr, 2016: The vertical structure of ocean eddies. Dyn. Stat. Climate Syst., 1, dzw001, https://doi.org/10.1093/climsys/dzw001.

    • Search Google Scholar
    • Export Citation
  • Deser, C., and J. M. Wallace, 1990: Large-scale atmospheric circulation features of warm and cold episodes in the tropical Pacific. J. Climate, 3, 12541281, https://doi.org/10.1175/1520-0442(1990)003<1254:LSACFO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Dewitte, B., and G. Reverdin, 2000: Vertically propagating annual and interannual variability in an OGCM simulation of the tropical Pacific Ocean in 1985–94. J. Phys. Oceanogr., 30, 15621581, https://doi.org/10.1175/1520-0485(2000)030<1562:VPAAIV>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Dottori, M., and A. J. Clarke, 2009: Rossby waves and the interannual and interdecadal variability of temperature and salinity off California. J. Phys. Oceanogr., 39, 25432561, https://doi.org/10.1175/2009JPO3898.1.

    • Search Google Scholar
    • Export Citation
  • Ebisuzaki, W., 1997: A method to estimate the statistical significance of a correlation when the data are serially correlated. J. Climate, 10, 21472153, https://doi.org/10.1175/1520-0442(1997)010<2147:AMTETS>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Enfield, D. B., and J. S. Allen, 1980: On the structure and dynamics of monthly mean sea level anomalies along the Pacific coast of North and South America. J. Phys. Oceanogr., 10, 557578, https://doi.org/10.1175/1520-0485(1980)010<0557:OTSADO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Frischknecht, M., M. Münnich, and N. Gruber, 2015: Remote versus local influence of ENSO on the California Current System. J. Geophys. Res. Oceans, 120, 13531374, https://doi.org/10.1002/2014JC010531.

    • Search Google Scholar
    • Export Citation
  • Fu, L.-L., and B. Qiu, 2002: Low-frequency variability of the North Pacific Ocean: The role of boundary- and wind-driven baroclinic Rossby waves. J. Geophys. Res., 107, 3220, https://doi.org/10.1029/2001JC001131.

    • Search Google Scholar
    • Export Citation
  • Gill, A. E., and A. J. Clarke, 1974: Wind-induced upwelling, coastal currents and sea-level changes. Deep-Sea Res. Oceanogr. Abtsr., 21, 325345, https://doi.org/10.1016/0011-7471(74)90038-2.

    • Search Google Scholar
    • Export Citation
  • Giunipero, E., and A. J. Clarke, 2013: Estimation of the effect of eddies on coastal El Niño flows using along-track satellite altimeter data. J. Phys. Oceanogr., 43, 12091224, https://doi.org/10.1175/JPO-D-12-0109.1.

    • Search Google Scholar
    • Export Citation
  • Grimshaw, R., and J. S. Allen, 1988: Low-frequency baroclinic waves off coastal boundaries. J. Phys. Oceanogr., 18, 11241143, https://doi.org/10.1175/1520-0485(1988)018<1124:LFBWOC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Herrera Cervantes, H., and A. Parés-Sierra, 1994: Propagación de variaciones de baja frecuencia en la temperatura superficial del Pacífico nor-oriental. Geofis. Int., 33, 469486, https://doi.org/10.22201/igeof.00167169p.1994.33.3.1184.

    • Search Google Scholar
    • Export Citation
  • Kessler, W. S., 1990: Observations of long Rossby waves in the northern tropical Pacific. J. Geophys. Res., 95, 51835217, https://doi.org/10.1029/JC095iC04p05183.

    • Search Google Scholar
    • Export Citation
  • Killworth, P. D., and J. R. Blundell, 1999: The effects of bottom topography on the speed of long extratropical planetary waves. J. Phys. Oceanogr., 29, 26892710, https://doi.org/10.1175/1520-0485(1999)029<2689:TEOBTO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Killworth, P. D., and J. R. Blundell, 2003: Long extratropical planetary wave propagation in the presence of slowly varying mean flow and bottom topography. Part I: The local problem. J. Phys. Oceanogr., 33, 784801, https://doi.org/10.1175/1520-0485(2003)33<784:LEPWPI>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Killworth, P. D., D. B. Chelton, and R. A. de Szoeke, 1997: The speed of observed and theoretical long extratropical planetary waves. J. Phys. Oceanogr., 27, 19461966, https://doi.org/10.1175/1520-0485(1997)027<1946:TSOOAT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Kutsuwada, K., and M. McPhaden, 2002: Intraseasonal variations in the upper equatorial Pacific Ocean prior to and during the 1997–98 El Niño. J. Phys. Oceanogr., 32, 11331149, https://doi.org/10.1175/1520-0485(2002)032<1133:IVITUE>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • LaCasce, J. H., 2017: The prevalence of oceanic surface modes. Geophys. Res. Lett., 44, 11 09711 105, https://doi.org/10.1002/2017GL075430.

    • Search Google Scholar
    • Export Citation
  • LaCasce, J. H., and S. Groeskamp, 2020: Baroclinic modes over rough bathymetry and the surface deformation radius. J. Phys. Oceanogr., 50, 28352847, https://doi.org/10.1175/JPO-D-20-0055.1.

    • Search Google Scholar
    • Export Citation
  • Li, J., and A. J. Clarke, 2004: Coastline direction, interannual flow and the strong El Niño currents along Australia’s nearly zonal southern coast. J. Phys. Oceanogr., 34, 23732381, https://doi.org/10.1175/JPO2645.1.

    • Search Google Scholar
    • Export Citation
  • Locarnini, R. A., and Coauthors, 2013: Temperature. Vol. 1, World Ocean Atlas 2013, NOAA Atlas NESDIS 73, 40 pp., http://data.nodc.noaa.gov/woa/WOA13/DOC/woa13_vol1.pdf.

  • McCreary, J. P., 1977: Eastern ocean response to changing wind systems. Ph.D. dissertation, Scripps Institution of Oceanography, University of California, 156 pp.

  • McCreary, J. P., 1984: Equatorial beams. J. Mar. Res., 42, 395430, https://doi.org/10.1357/002224084788502792.

  • Ramos, M., B. Dewitte, O. Pizzaro, and G. Garric, 2008: Vertical propagation of extratropical Rossby waves during the 1997–1998 El Niño off the west coast of South America in a medium-resolution OGCM simulation. J. Geophys. Res., 113, C08041, https://doi.org/10.1029/2007JC004681.

    • Search Google Scholar
    • Export Citation
  • Ramp, S. R., J. L. McClean, C. A. Collins, A. J. Semtner, and K. A. S. Hays, 1997: Observations and modeling of the 1991–1992 El Niño signal off central California. J. Geophys. Res., 102, 55535582, https://doi.org/10.1029/96JC03050.

    • Search Google Scholar
    • Export Citation
  • Rydbeck, A. V., T. G. Jensen, and M. Flatau, 2019: Characterization of intraseasonal kelvin waves in the equatorial Pacific Ocean. J. Geophys. Res. Oceans, 124, 20282053, https://doi.org/10.1029/2018JC014838.

    • Search Google Scholar
    • Export Citation
  • Samelson, R. M., 1992: Surface-intensified Rossby waves over rough topography. J. Mar. Res., 50, 367384, https://doi.org/10.1357/002224092784797593.

    • Search Google Scholar
    • Export Citation
  • Schopf, P. S., D. L. T. Anderson, and R. Smith, 1981: Beta dispersion of low-frequency Rossby waves. Dyn. Atmos. Oceans, 5, 187214, https://doi.org/10.1016/0377-0265(81)90011-7.

    • Search Google Scholar
    • Export Citation
  • Strub, P. T., and C. James, 2002: The 1997–1998 oceanic El Niño signal along the southeast and northeast Pacific boundaries—An altimetric view. Prog. Oceanogr., 54, 439458, https://doi.org/10.1016/S0079-6611(02)00063-0.

    • Search Google Scholar
    • Export Citation
  • Tailleux, R., and J. C. McWilliams, 2001: The effect of bottom pressure decoupling on the speed of extratropical, baroclinic Rossby waves. J. Phys. Oceanogr., 31, 14611476, https://doi.org/10.1175/1520-0485(2001)031<1461:TEOBPD>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Trenberth, K. E., 1984: Signal versus noise in the Southern Oscillation. Mon. Wea. Rev., 112, 326332, https://doi.org/10.1175/1520-0493(1984)112<0326:SVNITS>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Vega, A., Y. duPenhoat, B. Dewitte, and O. Pizzaro, 2003: Equatorial forcing of interannual Rossby waves in the eastern South Pacific. Geophys. Res. Lett., 30, 1197, https://doi.org/10.1029/2002GL015886.

    • Search Google Scholar
    • Export Citation
  • Vergara, O., B. Dewitte, M. Ramos, and O. Pizzaro, 2017: Vertical energy flux at ENSO time scales in the subthermoclinie of the southeastern Pacific. J. Geophys. Res. Oceans, 122, 60116038, https://doi.org/10.1002/2016JC012614.

    • Search Google Scholar
    • Export Citation
  • Wortham, C., and C. Wunsch, 2014: A multidimensional spectral description of ocean variability. J. Phys. Oceanogr., 44, 944966, https://doi.org/10.1175/JPO-D-13-0113.1.

    • Search Google Scholar
    • Export Citation
  • Wunsch, C., 1997: The vertical partition of oceanic horizontal kinetic energy. J. Phys. Oceanogr., 27, 17701794, https://doi.org/10.1175/1520-0485(1997)027<1770:TVPOOH>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Zhang, X., and A. J. Clarke, 2015: Observations of interannual equatorial freshwater jets in the western Pacific. J. Phys. Oceanogr., 45, 28482865, https://doi.org/10.1175/JPO-D-14-0245.1.

    • Search Google Scholar
    • Export Citation
  • Zhang, X., and A. J. Clarke, 2017: On the dynamical relationship between equatorial Pacific surface currents, zonally averaged equatorial sea level, and El Niño predictions. J. Phys. Oceanogr., 47, 323337, https://doi.org/10.1175/JPO-D-16-0193.1.

    • Search Google Scholar
    • Export Citation
  • Zlotnicki, V., Z. Qu, and J. Willis, 2019: MEaSUREs gridded sea surface height anomalies version 1812. PO.DAAC, accessed 22 May 2019, https://doi.org/10.5067/SLREF-CDRV2.

  • Zweng, M. M., and Coauthors, 2013: Salinity. Vol. 2, World Ocean Atlas 2013, NOAA Atlas NESDIS 74, 39 pp., http://data.nodc.noaa.gov/woa/WOA13/DOC/woa13_vol2.pdf.

  • Fig. 1.

    Contour plot of the local first baroclinic mode Rossby wave speed in deep waters (>3 km) under the conditions (a) w′ = 0 at z = −H, (b) w′ = 0 at z = −H with u¯ (see Killworth et al. 1997), (c) p′ = 0 at z = −H, and (d) p′ = 0 at z = −H with u¯. The Rossby wave speeds (cm s−1) are positive for westward propagation.

  • Fig. 2.

    Theoretical Rossby wave speeds for the first baroclinic mode for (a) the exclusion (“Standard”) and (b) inclusion (“Shallower Lags Included”) of grouped lags for the first 260 km from the coast. In each panel, zonal distance is regressed on theoretical lags grouped by 130-km zonal distance increments away from the California coast based on either bottom boundary condition w′ = 0 and p′ = 0 with and without baroclinic zonal mean flow inclusion (Killworth et al. 1997). Vertical error bars encompass the 130-km zonal distance per increment, while horizontal error bars are a standard deviation of the lags used to average within each increment. Note: the 95% confidence intervals found in Table 1 capture the variability depicted by the error bars on the wave speed estimates.

  • Fig. 3.

    Averaged zonal geostrophic velocity in the CCS (27°–37°N, 115°–125°W) based on WOA13-V2 climatological data for the years 1955–2012 for depths greater than or equal to 3 km.

  • Fig. 4.

    Monthly time series for interannual anomalies for coastal sea level measurements at San Diego, California (blue line; 32.7°N, 117.2°W), Santa Cruz, Ecuador (green line; 0.6°S, 90.3°W), and interannual anomalies of the Niño-3.4 SST index (averaged over the box 5°S–5°N, 170°–120°W). Each time series is created by removing the seasonal cycle and filtered using an 11-point symmetric interannual low-pass filter (Trenberth 1984). Each time series is continuous from January 1970 to December 2016 except for the Santa Cruz station where infrequent sampling becomes commonplace after May 2003. The maximum correlation between every time series occurs at lag-zero and is r = 0.72 [rcrit(99%) = 0.48] between San Diego vs Santa Cruz sea levels, r = 0.75 [rcrit(99%) = 0.42] for San Diego sea level vs Niño-3.4 SSTs, and r = 0.74 [rcrit(99%) = 0.49] for Santa Cruz sea level vs Niño-3.4 SSTs. All critical correlation coefficients, here and elsewhere, are computed using Ebisuzaki’s (1997) nonparametric statistical significance test.

  • Fig. 5.

    (top) Hovmöller time series of filtered (Trenberth 1984) interannual JPL MEaSUREs (Zlotnicki et al. 2019) SSH anomalies (cm) zonally averaged 1° off the eastern Pacific boundary for the latitudinal range 35°S–35°N. Our zonal average was taken to be 1° off the coast to capture the boundary signal. (middle) The primary EOF structural function of the zonally averaged interannual JPL SSH with percentage of explained variance (PEV = 86%). (bottom) Time series of the first principal component (blue line) and Niño-3.4 (orange line) with a maximum autocorrelation at lag 0 at r = 0.81 [rcrit(99%) = 0.52]. An interannual 11-point symmetric low-pass filter (Trenberth 1984) was utilized to capture the dominant interannual signal. This EOF analysis is an extension of Clarke (1992) who used 4 coastal stations from 15°S to 10°N to argue coastal pressure is spatially constant at low frequencies.

  • Fig. 6.

    (top) Regression of anomalous July–November surface wind (arrows) and outgoing longwave radiation (OLR) onto an equatorial SST anomaly index (anomalous July–November SST averaged over 6°S–6°N, 180°–80°W). Wind vectors are only shown for those grid points whose u or υ correlations with the SST index exceed 0.4 in absolute value. The OLR contour interval is 10 W m−2 °C−1 of the SST index. The zero contour is darkened, and the positive contour is dashed. Values less than −20 W m−2 °C−1 are shaded. The wind regressions are based on the period 1946–85 and the OLR regression on 1974–89 (1978 missing). (bottom) As in the top panel, but for December–February [redrawn from Deser and Wallace (1990)].

  • Fig. 7.

    First vertical EOF structure functions for the interannual pressure anomalies p′(z, t) in the equatorial Pacific at 96° and 84°W from January 1996 to July 2012. Also shown are the theoretical first vertical mode structure functions for the p′ = 0 (red) and the w′ = 0 (green) bottom boundary conditions. These theoretical modes were computed using WOA13-V2 climatological data. All plots are nondimensional and the theoretical modes are a best fit to the EOF structure (see text).

  • Fig. 8.

    Integrated EKW WKB ray paths beginning at the climatological mixed layer depth of 50 m at 150°W (de Boyer Montégut et al. 2004; Zhang and Clarke 2015). Shaded contours in the background represent the climatological buoyancy frequency as a function of depth and longitude averaged across the latitude band 2°S–2°N. The longitudinal domain extends from the eastern edge of interannual wind forcing (150°W) in the equatorial Pacific associated with El Niño events (see Fig. 6, top panel) to its eastern boundary (80°W). WKB ray paths reveal the depth of penetration for EKWs, whether forced in the interior or emitted from the western boundary, with dependence on angular frequency: ω = 2π/2 years (solid white line) and ω = 2π/3 years (dotted white line).

All Time Past Year Past 30 Days
Abstract Views 2398 2074 0
Full Text Views 958 746 150
PDF Downloads 314 138 14