1. Introduction
The El Niño–Southern Oscillation (ENSO) is among the most pervasive natural climate oscillations on earth, affecting the web of life from plankton to people. During mature El Niño (La Niña) events, the sea surface temperature (SST) in the eastern equatorial Pacific warms (cools), leading to global-scale responses in the terrestrial biosphere transmitted through modifications of large-scale atmospheric circulation. The dynamics of—and global responses to—ENSO have been studied for nearly eight decades (Walker and Bliss 1932; Ropelewski and Halpert 1989; Kiladis and Diaz 1989; Yulaeva and Wallace 1994). Cyclic patterns in climate events have also been connected to something resembling ENSO as early as the mid-nineteenth century. Reminiscing on his 1832 visit to Argentina during his expedition on the H.M.S. Beagle, British naturalist Charles Darwin notes “[t]hese droughts to a certain degree seem to be almost periodical; I was told the dates of several others, and the intervals were about fifteen years” (Darwin 1839). Nearly 60 years later, Darwin enters into his journal “. . . variations in climate sometimes appear to be the effect of the operation of some very general cause” (Darwin 1896). Some believe this “very general cause” was actually an early piecing together of ENSO and its now notorious impact on extreme weather events in South America (Cerveny 2005). It is only a coincidence that Darwin may have been among the first to point out the cyclic nature of ENSO, and the focus of this paper is the association between ENSO and the Galápagos Islands, which also owe their fame to Darwin.
Two fundamental characteristics of ENSO are amplitude, or the magnitude of the peak SST anomaly in the eastern equatorial Pacific, and period, or the time scale of the oscillation itself. These are also common metrics for evaluating the realism of interannual variability in coupled climate models (e.g., Collins 2000; Collins et al. 2006; Wittenberg et al. 2006). In principle, the atmospheric response to an ENSO event should be, to some extent, dependent on its amplitude. History will long remember the 1982–83 El Niño and the 1988 La Niña because they were strong, and the impacts were felt as such around the globe. One feature of the equatorial circulation that has not been considered with respect to ENSO amplitude and period are the Galápagos Islands. The Galápagos Archipelago is a group of 19 islands roughly 1000 km off the Ecuadorian coast of South America. Most of the islands of the Archipelago are miniscule, except for the largest, Isla Isabela, which accounts for 73% of the total land area of the Galápagos. Amounting to 5825 km2, Isla Isabela is larger than the U.S. state of Delaware. Given their prime location in the eastern equatorial Pacific Ocean, the Galápagos Islands are a popular location for the paleo community to extract time information on climate signals, including ENSO, from archives such as tropical corals. Here we consider whether the Galápagos Islands themselves are an important factor in the observed ENSO signal. The suggestion that the Galápagos affects ENSO may seem somewhat far fetched prior to examining its location. A vertical cross section of Galápagos topography is presented in Fig. 1. The cross section cuts across the longest possible north–south section through Isla Isabela, which is what is most relevant to the primarily zonal subsurface equatorial circulation. Isla Isabela covers completely the first degree south of the equator and crosses well into the first degree north of the equator.
The existence of the Galápagos directly on the equator presents the possibility of direct topographical interaction with the equatorial current system and other processes. Using an ocean general circulation model (OGCM) of the tropical Pacific Ocean, it was recently shown by Karnauskas et al. (2007) that inclusion of the Galápagos leads to an improved SST field. The implementation of the Galápagos into the ocean model used in Karnauskas et al. is also shown in Fig. 1. Breaking from the norm by northward displacing the “equatorial” gridpoint allows increased accuracy with respect to the actual dimensions of the island. The result was a warming in the east-central equatorial Pacific (spanning roughly 180°–90°W) up to 1.5°C on the annual mean. Currently, many ocean and coupled models, including that used in the NOAA Climate Forecast System (CFS), do not include the Galápagos. Despite the importance of the Pacific cold tongue as the location where ENSO events primarily manifest themselves in the surface ocean, many of these ocean and coupled models are also plagued by a cold bias, that is, the cold tongue is too cold, and a “double ITCZ,” which is a mysterious reduction in cloud cover over the equator, leaving symmetric cloud structures north and south of the equator (Stockdale et al. 1998; Harrison et al. 2002).
The equatorial undercurrent (EUC) is a critical component in the equatorial mass balance, and its obstruction by the Galápagos is critical to the improved SST, ultimately leading to a deeper and more diffuse thermocline, reduced entrainment/mixing, and reduced transport in the meridional vertical cells. The feedback between SST and net surface heat flux also contributes to the warming, which is possible because the ocean model used in Karnauskas et al. (2007) was coupled to an atmospheric mixed layer model such that heat fluxes were not prescribed. The pattern of SST warming due to the inclusion of the Galápagos is nearly identical to that of the known cold bias in current research and operational ocean models.
The next logical step is to understand what effect the Galápagos Islands have on interannual variability in the Pacific Ocean, that is, ENSO. There could also be important implications for the equatorial and coastal biological response to ENSO variability. The present is a follow-on study to Karnauskas et al. (2007) aimed at quantifying and explaining the effect of the Galápagos Islands on ENSO amplitude and period. The following section describes the models and methodology. Section 3 is a presentation and discussion of the results, and section 4 is a summary.
2. Models and datasets
a. Idealized forced experiments
The ocean model used in this study is identical to that used in Karnauskas et al. (2007): the Gent and Cane (1989) reduced gravity, primitive equation model of the tropical oceans with a hybrid vertical mixing scheme (Chen et al. 1994). The Chen et al. (1994) scheme accounts for mixed layer entrainment/detrainment, shear flow instability, and free convection in the thermocline by combining the physics of the Kraus and Turner (1967) mixed layer model with the Price et al. (1986) dynamical instability model. Surface fluxes are calculated interactively by coupling the OGCM to a thermodynamic atmospheric mixed layer (Murtugudde et al. 1996), thus allowing for feedbacks between SST and surface fluxes. The OGCM is structured vertically on sigma coordinates and includes a mixed layer plus 19 subsurface layers. The zonal boundaries of the model grid are 30°N and 20°S, along which a sponge layer is used (see Chen et al. 1994). Meridional boundaries are represented by the approximate coastlines of Asia, Indonesia, and Australia in the western pacific (Indonesian Throughflow is closed off) and the Americas in the eastern Pacific. All experiments are run on a grid with higher resolution in the eastern tropical Pacific: zonally stretching from ¼° in the east to ∼1° at the western boundary, and meridional stretching from ¼° near the equator to ∼1° at 20°S and 30°N. Since the OGCM is a reduced gravity model, all topography is constant in the vertical. Thus, there are no realistic continental shelves, implying that the experiments including the Galápagos essentially represent the islands as a simple vertical barrier extending from the ocean floor to the sea surface. However, based on subsurface bathymetric measurements (Fig. 1), this is a reasonable approximation in the upper few hundred meters of the ocean, where most of the interesting features of the equatorial circulation are at play.
In the forced setup, two experiments were performed—designed to reveal the total effect of the Galápagos Islands on an idealized El Niño event. Prior to both experiments, the model was spun up from the climatology of Levitus and Boyer (1994) for 60 years. The first of the two forced experiments, NoGF, was initialized with conditions that were produced by spinning up the model with the Galápagos excluded from the model grid. Thus, the mean state in NoGF is the mean state that would exist if the Galápagos Islands did not exist. The model was then integrated for two years, saving 10-day mean fields. The second of the two forced experiments, GF, was initialized with initial conditions produced by spinning up the model with the Galápagos Islands included in the model grid. Thus, the mean state in GF is the appropriate mean state accounting for the existence of the Galápagos. As mentioned in section 1 and explained in detail in Karnauskas et al. (2007), the Galápagos mean state contains a reduction of the equatorial cold bias over the non-Galápagos mean state. The model was then integrated for two years, saving 10-day mean fields.
Each forced experiment (e.g., NoGF) comprises two model runs: one forced by climatological forcing and one forced by climatological forcing plus a westerly wind stress anomaly designed to induce a reasonable El Niño event. Thus, anomalies for each experiment are computed as the difference between the perturbation and climatological runs of that experiment. A westerly wind stress anomaly of 0.072 N m−2 (equivalent to about 6 m s−1) was added to the climatological zonal wind stress over a box centered on the equator in the western Pacific (maximum at the equator tapered outward to 8°N–8°S, beween 160°E and 160°W) from September through November of the first year of integration. The location, magnitude, timing, and duration of the idealized wind stress anomaly are reasonably similar to those preceding significant El Niño events in recent observed history.
Climatological forcing is derived from ECMWF operational analysis surface wind stress, 2° spatial resolution, base period 1985–2003 (Bengtsson et al. 1982); Xie and Arkin (1996) precipitation, 2.5° spatial resolution, base period 1979–2003; International Satellite Cloud Climatology Project (ISCCP) cloud cover, 280-km spatial resolution, base period 1983–94 (Rossow and Schiffer 1991); and Earth Radiation Budget Experiment (ERBE) shortwave radiation, 2.5° spatial resolution, base period 1984–90 (Barkstrom 1984).
Two distinct types of ENSO modes, which can be identified by the direction of propagation of the SST anomaly, are the SST mode (or S mode), which exhibits westward propagation and involves local SST–wind interactions, and the thermocline mode (or T mode), which exhibits eastward propagation and involves remote wind–thermocline interactions (Guilyardi 2006). In that light, the forced El Niño events analyzed in this study are of the T-mode variety, although it is understood that ENSO events in nature are not as straightforward. Since the purpose of this study is, not to understand the genesis of ENSO events but, to quantify and understand the effect of the Galápagos Islands on key ENSO characteristics, we suggest that the method used to initiate El Niño events in the forced experiments is of secondary importance. Furthermore, the amplitude of the idealized El Niño event produced by the model is arbitrary. The greater the westerly wind stress anomaly we impose, the higher the amplitude of the ensuing El Niño event. Again, the goal is, not to accurately simulate a specific El Niño event as seen in observations but, rather, to study characteristics that are dependent on the Galápagos Islands.
b. Hybrid coupled experiments


Two 75-yr simulations were produced, with the sole difference being the existence of the Galápagos Islands in the model grid. Experiment NoGC was initialized with the spunup state without Galápagos (the same initial conditions as those used in idealized forced experiment NoGF) and integrated forward with the island absent from the model grid. Experiment GC was initialized with the spunup state that included the Galápagos Islands (the same initial conditions as those used in idealized forced experiment GF) and integrated forward with the Galápagos present in the model grid. After removing the brief spinup period due to the introduction to the SST–wind stress coupling, 68 years of output are available for analysis.
3. Results
a. Idealized forced experiments
In this brief subsection, we examine the effect of the Galápagos Islands on an idealized El Niño event and the mechanisms responsible for the differences. We first examine indices that highlight the ocean response at key locations in the tropical Pacific Ocean. Niño-4 (5°S–5°N, 160°E–150°W) represents the anomaly in the west-central equatorial Pacific, Niño-3 (5°S–5°N, 150°–90°W) in the eastern Pacific, and Niño-1 + 2 (10°S–0°, 90°–80°W) in the coastal upwelling region off the west coast of South America but east of the Galápagos Islands.
Area-averaged SST anomalies were computed for each of the aforementioned Niño indices and are presented in Fig. 2. According to Fig. 2, the peak SST anomaly in the Niño-3 region is reduced by approximately 0.5°C by the Galápagos, or a 15% reduction in the anomaly over experiment NoGF. The reduction of the anomaly in Niño-4 is small. In Niño-1 + 2, however, the SST anomaly is reduced by approximately half. The lag between the three Niño indices is relatively unchanged between NoGF and GF, with both experiments exhibiting Niño-4 leading, followed by Niño-3, and finally Niño-1 + 2.


b. Hybrid coupled experiments
In this section, we discuss the effect of the Galápagos Islands on internal ENSO variability with the ocean model coupled to the atmosphere through zonal wind stress. Specifically, we are interested in any changes in the time scale of ENSO, as the idealized experiments served to show that, given an identical perturbation in the wind field, the amplitude of the SST response should be damped in the case with the Galápagos Islands. Recall that the sole difference between hybrid coupled experiments NoGC and GC is the existence of the island in the ocean model grid. It is well known that ENSO events tend to peak in boreal winter, with Niño-3 SST anomaly (SSTA) variance peaking strongly in November and December. In this regard, both the NoGC and GC experiments are in excellent agreement with observations (not shown).
The Niño-3 SST time series from both experiments, with the seasonal cycle removed, are shown in Fig. 4 (top). Also shown in Fig. 4 (bottom) are the power spectra computed by discrete Fourier transform (DFT) for Niño-3 SSTA in experiments NoGC, GC, and 68 years (1935–2002) of the NOAA Extended Reconstructed SST version 2 dataset (Smith and Reynolds 2004). To avoid spurious spectral features (i.e., those of period close to the length of the time series), we only show power spectra from the fourth harmonic and beyond (i.e., period not longer than one quarter the length of the time series). Examining the power spectrum for the NoGC experiment, there is one dominant spectral peak, which is found at the 2-yr period. This can be confirmed intuitively by counting the number of warm events in any 10-yr interval of the NoGC Niño-3 time series; there will be exactly 5. This is classic “biennial ENSO” behavior exhibited by some ocean and coupled climate models. Such strictly biennial regularity is not realistic, as ENSO in nature at present is neither perfectly regular nor significantly biennial. In our coupled model simulations, the Galápagos Islands have the effect of significantly reducing the biennial peak and giving rise to a preferred time scale near 3 yr. Moreover, a closer look at the GC Niño-3 time series reveals that the system also has the ability to transition into and out of a biennial ENSO regime; between years 20 and 40, the ENSO periodicity in the GC experiment is nearly biennial. It is also interesting to note that during, that period, the amplitude is smaller than surrounding periods. In NoGC, ENSO is strictly biennial throughout with relatively consistent amplitude.
As evident by comparing simulated and observed power spectra (Fig. 4, bottom), the overall effect of reducing the biennial peak and increasing spectral power at lower interannual frequencies represents considerable improvement in the representation of ENSO, even within a relatively simple coupled context. An alternative technique for computing power spectra, which readily lends itself to understanding the statistical significance of spectral peaks, is the multitaper method (MTM) (Ghil et al. 2002). From the MTM method, the basic features of the NoGC and GC Niño-3 power spectra are very similar to those resulting from DFT (Fig. 4). In the NoGC Niño-3 power spectrum, the 2-yr peak is the only interannual peak significant at the 99% confidence level. In the GC power spectrum, a broad swath of three peaks (from ∼3 to 1.5 yr) is significant at the 99% confidence level.
To examine why the Galápagos Islands leads to a shift in the ENSO time scale from biennial to quasi quadrennial, we begin by analyzing composite ENSO events. We perform an empirical orthogonal function analysis on model SST for the 68 years of output, which is presented in Fig. 5. The only real mode possible in our free-running hybrid coupled model is ENSO, as there is no interannual forcing being applied to the system that would justify otherwise. The first EOF of SSTA in both NoGC and GC experiments represents a fully mature ENSO event, portrayed in its warm phase in Fig. 5. This is evident by the fact that the first principal component (PC1) and the Niño-3 index in each experiment are correlated >0.99 for years 42–68 (Fig. 5, bottom left). Despite the requirement of orthogonality, an EOF analysis can lead to “propagating” modes, that is, individual statistical modes representing different aspects or phases of a single phenomenon. The second (and third, in the case of GC) mode of SSTA represents the transition between positive and negative phases of ENSO, that is, the warming period preceding an El Niño event and cooling that follows. The warming and cooling periods surrounding mature ENSO events are akin to the process described as recharging and discharging heat from the equatorial Pacific (e.g., Jin 1997); these secondary modes of SSTA variability could be said to correspond with the recharge–discharge process. Spatially, there are only minor differences in SST modes between experiments NoGC and GC. The exception is a break in the EOF loading at the western edge of the Galápagos Islands, which is found farther toward the coast of South America in the NoGC case.
Temporally, differences in the EOF analysis shed light on an important distinction in the nature of ENSO in the coupled experiments. In both experiments, the secondary principal components (PC2) are maximum while the PC1/Niño-3 signal is growing, and minimum while the PC1/Niño-3 signal is decaying (again, because they represent the warming and cooling phases). However, in the GC experiment, the positive phases in the PC2 are smaller in magnitude and longer lasting, consistent with a slower growth of PC1/Niño-3 signal. Conversely, the PC2 signal in the NoGC experiment grows as rapidly as it decays, and does not persist in either phase. Hence, the ENSO events themselves in the NoGC experiment grow and decay similarly, resulting in a shorter period than in the case with the Galápagos Islands. The present discussion is equally applicable to model sea level anomaly (SLA), as in Fig. 6. The total SLA is subject to the effects of propagating Kelvin and Rossby waves characteristic of the delayed oscillator paradigm of ENSO, and the heat content changes associated with the recharge–discharge process of ENSO, particularly in the eastern equatorial Pacific where the broader thermocline variations are large (e.g., McPhaden and Yu 1999). As with SSTA, the leading PC of SLA corresponds with the phase of ENSO in both experiments (Fig. 6, bottom), but the warming phase is slower evolving in the GC experiment.
To examine the differences in the time evolution of composite ENSO events between the two hybrid coupled experiments, Fig. 7 portrays the composite (over years 42–68) evolution of Niño-3 SSTA, SLA PC, SSTA PCs, as well as the time rate of change of the Niño-3 mixed layer heat content anomaly d(HC′ml)/dt and the zonal wind stress anomaly averaged from 5°S to 5°N. The basic message is that Niño-3 SSTA, the sea level pattern, SST PC1, and zonal wind stress anomaly track very closely but grow much more rapidly in the NoGC experiment. This is consistent with the fact that the maxima in the secondary SSTA mode, which represents the warming preceding El Niño events, and the change in mixed layer heat content itself, are shorter lived in the NoGC experiment and essentially mirror the minima following the event. In the GC case, the warming mode persists longer and changes the mixed layer heat content more gradually (i.e., a smaller rate of change over a longer period of time). A direct comparison of the composite SST anomaly in the eastern equatorial Pacific is shown in Fig. 7c and reveals a slower growth but, ultimately, the same amplitude. When this slower and more subtle growth of the warm anomaly is compared to a composite of observed El Niño events (Fig. 7d), it is clearly an improvement in the representation of ENSO evolution. The mechanisms behind the slower growth of warm ENSO events in the hybrid coupled model, which lead to a longer preferred time scale for ENSO, and how they relate the Galápagos Islands are discussed in the following subsection.
c. The delayed oscillator and Bjerknes feedback
In this subsection, we discuss the mechanism for the change in ENSO time scale. There are several important theories for the oscillatory nature of ENSO. The concept of the delayed oscillator (DO) mechanism as proposed by Suarez and Schopf (1988) implies that the ENSO period is set by the basin crossing times of the equatorially trapped oceanic waves. Strictly within the context of the DO mechanism and its constituent equatorial waves, there are two ways by which the introduction of the Galápagos Islands to the hybrid coupled model could lead to a change in the ENSO period: 1) the speed of the Kelvin or Rossby waves change, thereby changing the basin crossing times or 2) the Galápagos Islands changes the effective width of the ocean basin by acting as an effective eastern boundary, thereby changing the basin crossing times of the equatorial waves. The latter is not a plausible explanation for two reasons: While the Galápagos Islands are wide enough in the meridional direction to significantly obstruct the EUC and cause a shift in the mean state (Karnauskas et al. 2007), they are not wide enough to cause any appreciable reflection of the Kelvin wave. Second, if the Galápagos Islands were wide enough to act as an effective eastern boundary and reflect Kelvin waves, this would result in shorter basin crossing times for the Kelvin wave and translate into a shorter ENSO period—the opposite of which is the case.




Rather than explain the change in ENSO period as a result of the Galápagos Islands from the standpoint of the DO set by the wave speed the waves, we consider the behavior of the fully coupled system and its response to Kelvin waves under different mean states, particularly the role of the Bjerknes feedback (Bjerknes 1969), while invoking aspects of other leading theories for the oscillatory nature of ENSO, including the recharge–discharge mechanism (Jin 1997). From Eq. (3) and its solution, illustrated in Fig. 8, the solution for the zonal geostrophic current associated with an equatorial Kelvin wave is damped in the presence of a larger H, which is the case in the GC mean state. Whether an equatorial Kelvin wave is of the downwelling or upwelling variety is defined by the uKe field; a convergent (divergent) uKe field results in downwelling (upwelling) signal propagating along with the wave energy. Similarly, a weaker convergence of uKe imparts weaker downwelling and therefore results in a smaller SST anomaly. This, in part, helps to explain the damped SST anomaly response to identical wind stress perturbations in the idealized forced experiments. However, in the coupled context, this has further implications for the growth of the SST anomaly in time.
To analyze the behavior (amplitude and temporal characteristics) of Kelvin waves in the hybrid coupled model experiments, we compute the projection of Kelvin waves onto the geostrophic current anomaly. Figure 10 (top) shows the ratio of the variability of the Kelvin wave signal along the equator from 180° to 100°W in experiments NoGC and GC. Also shown in Fig. 10 (bottom) is the difference in the mean state (equatorial temperature) without the Galápagos Islands (annual mean). Because the interannual variability inherent to the coupled experiments would be aliased into any long-term mean calculations, we rely on the difference in annual means derived from the climatological runs of the idealized forced experiments. As expected from theoretical considerations (Fig. 8), the Kelvin wave variability is damped by up to 28% by the Galápagos, which appears to be stronger whereas the thermocline difference is greater (e.g., note similarities at 105° and 125°W). Note that the standard deviation is dependent only on amplitude and not frequency: thus, the result in Fig. 10 is independent of the fact that there are fewer full ENSO cycles in the GC case. East of the Galápagos Islands, the thermocline is much deeper in the Galápagos mean state. The Kelvin wave may therefore become more damped in that region, although this cannot be confirmed due to problems with calculating projections near a concave coastline.


d. Subtropical cells and recharge–discharge
With the rate of anomaly growth and thus time scale difference reconciled, the curiosity remains: Why, if the Kelvin wave SST response is damped, would the amplitude of the fully mature warm ENSO event be similar (e.g., the composite events shown in Fig. 7c)? As previously alluded, ENSO and its oscillatory nature have been described as a series of recharge and discharge phases (Jin 1997). The meridional circulation in the tropical–subtropical Pacific Ocean is defined by wind-driven subtropical cells, vigorous upwelling at the equator, poleward surface transport, subduction, and equatorward return flow along the thermocline. Variations in transport convergence in the upper pycnocline by Pacific subtropical cells have recently gained attention due to their possible role in modulating decadal variability (McPhaden and Zhang 2002, 2004). Greater transport convergence is associated with cooler equatorial SST, and vice versa, because the equatorward flow in the upper pycnocline is a branch in the same circuit as the equatorial upwelling. Likewise, suppressed transport convergence indicates reduced equatorial upwelling of cold water and would lead to a warmer equatorial SST. In this context, we examine the role of mass transport convergence in the upper pycnocline in affecting differences between the simulated ENSO in our hybrid coupled experiments. Our methodology is identical to that of McPhaden and Zhang (2002, 2004). Based on the vertical structure of the upper pycnocline described by McPhaden and Zhang and vertical variations of mass transport, in either hemisphere and for either coupled simulation, we choose density surfaces between which to vertically integrate mass transport. In experiment NoGC we use 21.18–25.19 kg m−3 (21.71–25.06 kg m−3) for the Northern (Southern) Hemisphere and in experiment GC, respectively, 21.16–24.85 kg m−3 (21.71–24.68 kg m−3). Finally, along 6°N (6°S), we integrate from 145°E–90°W (160°E–85°W) and obtain upper pycnocline convergence (UPC) by taking the difference between the northward transport at 6°S and the southward transport at 6°N.
We again construct a picture of the composite El Niño event in terms of UPC (Fig. 12). The unfolding of events in the NoGC composite is as follows: a large SSTA develops at an early stage because the mean state is one in which SSTA is highly sensitive to Kelvin waves, a strong Bjerknes feedback drives a rapid growth of both the SSTA and zonal wind stress anomaly (recall from Figs. 7a,b that the SSTA and zonal wind stress anomaly lines are nearly identical), and UPC is immediately and strongly suppressed. The UPC minimum occurs at the same time as the SSTA maximum. In effect, the DO and Bjerknes feedback are operating in close synchrony, and the UPC quickly responds to the large fluctuations in zonal wind stress. Over the 68 years simulated, the UPC anomaly and SSTA time series are anticorrelated −0.82. This is indicative of a strong positive feedback between UPC and SST during the warming phase: the SST warming reduces the zonal winds, which reduces UPC and upwelling, which contributes to the SST warming. Also in the NoGC composite (Fig. 12a), UPC is in quadrature with d(HC′ml)/dt; as UPC decreases, HC′ml increases.
The evolution of UPC in the GC composite is quite different: a smaller SSTA begins to develop because the mean state is one in which SSTA is less sensitive to Kelvin waves, and a weaker Bjerknes feedback drives a slower growth of the SSTA and zonal wind stress anomaly. That wind stress anomaly is also apparently too weak throughout most of the warming period to cause a change in the wind-driven UPC. As a result, UPC does not begin to decrease until later than in the NoGC case. At the end of the warming phase, the zonal wind stress anomaly is strong enough to suppress the UPC, which drives the Niño-3 SSTA to approximately the same amplitude as in the NoGC composite. As evident in the GC composite, there is strong association between UPC and d(HC′ml)/dt, meaning that UPC is having a negative effect on the warming throughout the composite event.
The recharge–discharge paradigm of ENSO was also discussed by Kessler (2002) who, using observations, portrayed the ENSO cycle as an evolution of the state of the equatorial system through phase space defined by basin-wide mean 20°C isotherm depth and Niño-3 SST. In a similar fashion, Fig. 13 depicts the coupled ENSO cycle as an evolution through phase space defined by Niño-3 SSTA and zonal wind stress anomaly (averaged from 5°N to 5°S). In the NoGC case (Fig. 13a), the system undergoes a rapid growth/warming (quadrant IV to quadrant I) and immediately undergoes a rapid decay/cooling (quadrant II to quadrant III), with little time spent near the origin. Conceptually, this is the DO mechanism operating on a more sensitive mean state with the Bjerknes feedback further driving the anomaly growth. In Kessler’s phase diagram depicting the recharge–discharge process, this would resemble a constant circular orbit through the four quadrants of phase space with nearly equal time spent in each state, rather than an abundance of points near the origin or in the recharge phase. In contrast and more consistent with observations, the GC phase diagram (Fig. 13b) depicts a DO mechanism operating on a less sensitive mean state, with the Bjerknes feedback acting more subtly until the equatorial system is completely discharged and promptly returns to gradually recharge (i.e., more data points are found near the origin and in quadrant III). Such an evolution through the coupled phase space is more realistic with respect to observations (Fig. 13c).
4. Summary and implications
In this paper, we have described the effect of the Galápagos Islands on forced El Niño events and coupled ENSO variability. The Galápagos result in a reduced SST anomaly in response to the forced El Niño throughout the equatorial Pacific, but especially east of the island and beneath the surface, which is primarily due to the different mean state, that is, a deeper and more diffuse thermocline.
The power spectrum of the hybrid coupled experiment without the Galápagos Islands is dominated by a peak at 2 yr, which is the classic biennial ENSO characteristic of many ocean and coupled climate models. Two such examples are the NCAR Community Climate System Model, version 3 (CCSM3) (Collins et al. 2006) and the ECHAM4 (Bacher et al. 1998; Guilyardi et al. 2004). Such strictly biennial regularity is not realistic, as ENSO in nature is neither perfectly regular nor significantly biennial. The Galápagos Islands have the effect of significantly reducing the biennial peak and generating a quasi-quadrennial preferred time scale, which closer matches observations. We explain the shift in ENSO period in the context of current paradigms for the oscillatory nature of ENSO. The delayed oscillator model of ENSO would dictate a shorter period of ENSO. However, given a damped SST anomaly, which is the overall effect of the Galápagos mean state, the growth of the SST anomaly through the Bjerknes feedback is slowed. A slower warming translates into a longer complete cycle as measured between crests or troughs in a time series of, say, the Niño-3 SST anomaly. Fedorov and Philander (2000) discussed the relationship between the mean state and the period of ENSO; our results are highly consistent with their notion that a deeper mean thermocline results in a longer ENSO period.
The Pacific subtropical cells also contribute to the slowness of the warming preceding the peak El Niño. Since the transport convergence is related to the strength of the equatorial winds, a gentle weakening of the trades, as is the case in the coupled experiment that included the Galápagos Islands, does not lead to suppression of the subsurface transport convergence and therefore does not contribute to the warming. The amplitude of the peak event in the GC experiments is approximately as large as in NoGC, however, because the transport convergence is eventually suppressed once the zonal wind stress anomaly is finally strong enough to affect a change in the subsurface meridional transport convergence. Others have argued that, within the recharge–discharge paradigm, the pattern of zonal wind stress variability may be a potential reason for global coupled models having an ENSO period shorter than observed (Capotondi et al. 2006). Given the coupled nature of ENSO, the time scale of ENSO is likely the combined result of multiple factors, as it appears to be in coupled models.
The most general implication is that ENSO frequency is highly sensitive to the mean state, and, if one is to properly simulate ENSO variability in even a simple coupled model, it is particularly crucial to accurately represent the thermocline depth and mixed layer physics. In our case and based on the spatial patterns of SST biases in other models, it is the introduction of the Galápagos Islands that seems to bring about the right thermocline change to enable important aspects of tropical ocean–atmosphere dynamics and, ultimately, a more realistic ENSO frequency. For models that already have a thermocline that is too deep or too diffuse, the Galápagos Islands may not lead to such improvements. There are examples of coupled models that are plagued by the usual equatorial cold bias, while exhibiting reasonable ENSO frequency, such as the Geophysical Fluid Dynamics Laboratory Climate Model version 2.0 (GFDL CM2.0/2.1) (Wittenberg et al. 2006). It has also been argued that some coupled GCMs produce a biennial ENSO because of an oversensitivity to interactions between the Pacific and Indian Oceans (Yu 2005). The Galápagos Islands were not included in the model grid of Yu (2005). We do not suggest that our results conflict with those of Yu (2005); strong Indo-Pacific interactions could be responsible for an excessively biennial ENSO in some coupled models. However, such interactions in models are sensitive to the treatment of the Indonesian Throughflow and the highly complex topography that acts as a partial barrier between the Indian and Pacific Oceans. Our model, as set up for the present experiments, does not include any form of Indo-Pacific interaction.
Such a change in the overall amplitude and time scale for ENSO, at least as contained within the Pacific basin, should have a considerable impact on the role of—and response to—ENSO in the broader global atmospheric circulation. In global models plagued by a strong biennial ENSO, our results suggest that the introduction of the Galápagos Islands may, in some cases, be a possible solution. This is not to say the Galápagos is the only possible solution; several studies have also shown that the atmospheric component of a coupled model can make a difference in ENSO characteristics (e.g., Schneider et al. 2003; Guilyardi et al. 2004). In terms of operational predictions of the coupled system, The NOAA Environmental Modeling Center (EMC) is currently assessing the impact of the Galápagos in the Climate Forecast System (CFS) (Jiande Wang 2007, personal communication). This also applies to physical–ecosystem modeling since the Galápagos Islands interact with Kelvin waves prior to propagating through the east Pacific warm pool and over Costa Rica Dome—a region of complex ocean biogeochemistry.
Acknowledgments
The authors thank Mr. Jim Beauchamp and Mr. Eric Hackert for assistance with managing model output and useful discussions. This research was supported by the National Oceanic and Atmospheric Administration (NOAA) Pan American Climate Studies (PACS) program through Grant NA17EC1483.
REFERENCES
Bacher, A., J. M. Oberhuber, and E. Roeckner, 1998: ENSO dynamics and seasonal cycle in the Tropical Pacific as simulated by the ECHAM4/OPYC3 coupled general circulation model. Climate Dyn., 14 , 431–450.
Barkstrom, B. R., 1984: The Earth Radiation Budget Experiment (ERBE). Bull. Amer. Meteor. Soc., 65 , 1170–1186.
Bengtsson, L., M. Kanamitsu, P. Kallberg, and S. Uppala, 1982: FGGE 4-dimensional data assimilation at ECMWF. Bull. Amer. Meteor. Soc., 63 , 29–43.
Bjerknes, J., 1969: Atmospheric teleconnections from the equatorial Pacific. Mon. Wea. Rev., 97 , 163–172.
Capotondi, A., A. Wittenberg, and S. Masina, 2006: Spatial and temporal structure of tropical Pacific interannual variability in the 20th century coupled simulations. Ocean Modell., 15 , 274–298.
Cerveny, R. S., 2005: Charles Darwin’s meteorological observations aboard the H.M.S. Beagle. Bull. Amer. Meteor. Soc., 86 , 1295–1301.
Chen, D., L. M. Rothstein, and A. J. Busalacchi, 1994: A hybrid vertical mixing scheme and its application to tropical ocean models. J. Phys. Oceanogr., 24 , 2156–2179.
Collins, M., 2000: The El Niño–Southern Oscillation in the second Hadley Centre coupled model and its response to greenhouse warming. J. Climate, 13 , 1299–1312.
Collins, W. D., and Coauthors, 2006: The Community Climate System Model version 3 (CCSM3). J. Climate, 19 , 2122–2143.
Darwin, C., 1839: Journal of Researches into the Geology and Natural History of the Various Countries Visited by H.M.S. Beagle. Hafner Publishing, 615 pp. (1952, reprint.).
Darwin, C., 1896: Journal of Researches into the Natural History and Geology of the Countries Visited during the Voyage of the. H.M.S. Beagle Round the World, under the Command of Capt. Fitz Roy, R.N. 2nd ed. D. Appleton and Company, 519 pp.
Fedorov, A. V., and S. G. H. Philander, 2000: Is El Niño changing? Science, 288 , 1997–2002.
Gent, P. R., and M. A. Cane, 1989: A reduced gravity, primitive equation model of the upper equatorial ocean. J. Comput. Phys., 81 , 444–480.
Ghil, M., and Coauthors, 2002: Advanced spectral methods for climatic time series. Rev. Geophys., 40 , 1–41.
Gill, A. E., 1982: Atmosphere–Ocean Dynamics. Academic Press, 662 pp.
Guilyardi, E., 2006: El Niño–mean state–seasonal cycle interactions in a multi-model ensemble. Climate Dyn., 26 , 329–348.
Guilyardi, E., and Coauthors, 2004: Representing El Niño in coupled ocean–atmosphere GCMs: The dominant role of the atmospheric component. J. Climate, 17 , 4623–4629.
Harrison, M. J., A. Rosati, B. J. Soden, E. Galanti, and E. Tziperman, 2002: An evaluation of air–sea flux products for ENSO simulation and prediction. Mon. Wea. Rev., 130 , 723–732.
Holton, J. R., 1992: An Introduction to Dynamic Meteorology. 3rd ed. Academic Press, 511 pp.
Jin, F. F., 1997: An equatorial ocean recharge paradigm for ENSO. Part I: Conceptual model. J. Atmos. Sci., 54 , 811–829.
Kalnay, E., and Coauthors, 1996: The NCEP/NCAR 40-Year Reanalysis Project. Bull. Amer. Meteor. Soc., 77 , 437–471.
Karnauskas, K. B., R. Murtugudde, and A. J. Busalacchi, 2007: The effect of the Galápagos Islands on the equatorial Pacific cold tongue. J. Phys. Oceanogr., 37 , 1266–1281.
Kessler, W. S., 2002: Is ENSO a cycle or a series of events? Geophys. Res. Lett., 29 .2125, doi:10.1029/2002GL015924.
Kiladis, G. N., and H. F. Diaz, 1989: Global climatic anomalies associated with extremes in the Southern Oscillation. J. Climate, 2 , 1069–1090.
Kraus, E. B., and S. J. Turner, 1967: A one-dimensional model of the seasonal thermocline. Part II. Tellus, 19 , 98–105.
Levitus, S., and T. Boyer, 1994: Temperature. Vol. 4. World Ocean Atlas 1994, NOAA Atlas NESDIS 4, 117 pp.
Matsuno, T., 1966: Quasi-geostrophic motions in the equatorial area. J. Meteor. Soc. Japan, 44 , 24–42.
McPhaden, M., and X. Yu, 1999: Equatorial waves and the 1997–1998 El Niño. Geophys. Res. Lett., 26 , 2961–2964.
McPhaden, M. J., and D. Zhang, 2002: Slowdown of the meridional overturning circulation in the upper Pacific Ocean. Nature, 415 , 603–608.
McPhaden, M. J., and D. Zhang, 2004: Pacific Ocean circulation rebounds. Geophys. Res. Lett., 31 .L18301, doi:10.1029/2004GL020727.
Murtugudde, R., R. Seager, and A. Busalacchi, 1996: Simulation of the tropical oceans with an ocean GCM coupled to an atmospheric mixed layer model. J. Climate, 9 , 1795–1815.
Price, J. F., R. A. Weller, and R. Pinkel, 1986: Diurnal cycling: Observations and models of the upper ocean response to diurnal heating, cooling, and wind mixing. J. Geophys. Res., 91 , 8411–8427.
Reynolds, R. W., N. A. Rayner, T. M. Smith, D. C. Stokes, and W. Wang, 2002: An improved in situ and satellite SST analysis for climate. J. Climate, 15 , 1609–1625.
Ropelewski, C. F., and M. S. Halpert, 1989: Precipitation patterns associated with the high index phase of the Southern Oscillation. J. Climate, 2 , 268–284.
Rossow, W. B., and R. A. Schiffer, 1991: ISCCP cloud data products. Bull. Amer. Meteor. Soc., 72 , 2–20.
Schneider, E. K., D. G. DeWitt, A. Rosati, B. P. Kirtman, L. Ji, and J. J. Tribbia, 2003: Retrospective ENSO Forecasts: Sensitivity to atmospheric model and ocean resolution. Mon. Wea. Rev., 131 , 3038–3060.
Smith, T. M., and R. W. Reynolds, 2004: Improved extended reconstruction of SST (1854–1997). J. Climate, 17 , 2466–2477.
Stockdale, T. N., A. J. Busalacchi, D. D. Harrison, and R. Seager, 1998: Ocean modeling for ENSO. J. Geophys. Res., 103 , 14325–14356.
Suarez, M. J., and P. S. Schopf, 1988: A delayed action oscillator for ENSO. J. Atmos. Sci., 45 , 3283–3287.
Sun, D-Z., T. Zhang, and S-I. Shin, 2004: The effect of subtropical cooling on the amplitude of ENSO: A numerical study. J. Climate, 17 , 3786–3798.
Walker, G. T., and E. W. Bliss, 1932: World weather V. Mem. Roy. Meteor. Soc., 4 , 53–84.
Wang, W., and M. J. McPhaden, 2000: The surface-layer heat balance in the equatorial Pacific Ocean. Part II: Interannual variability. J. Phys. Oceanogr., 30 , 2989–3008.
Wittenberg, A. T., A. Rosati, N-C. Lau, and J. J. Ploshay, 2006: GFDL’s CM2 global coupled climate models. Part III: Tropical Pacific climate and ENSO. J. Climate, 19 , 698–722.
Xie, P., and P. A. Arkin, 1996: Analyses of global monthly precipitation using gauge observations, satellite estimates, and numerical model predictions. J. Climate, 9 , 840–858.
Yu, J-Y., 2005: Enhancement of ENSO’s persistence barrier by biennial variability in a coupled atmosphere–ocean general circulation model. Geophys. Res. Lett., 32 .L13707, doi:10.1029/2005GL023406.
Yulaeva, E., and J. M. Wallace, 1994: The signature of ENSO in global temperature and precipitation fields derived from the microwave sounding unit. J. Climate, 7 , 1719–1736.

(left) Map of the Galápagos region with 1-km resolution, model grid lines (thin gray lines), and grid points used to represent the Galápagos Islands in the model (heavy black lines). (right) Vertical section of subsurface topography across Isla Isabela, constructed from multiple sources of bathymetric measurements along 91.25°W (black dashed line in left). Bathymetric data compiled by William Chadwick, Oregon State University
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

(left) Map of the Galápagos region with 1-km resolution, model grid lines (thin gray lines), and grid points used to represent the Galápagos Islands in the model (heavy black lines). (right) Vertical section of subsurface topography across Isla Isabela, constructed from multiple sources of bathymetric measurements along 91.25°W (black dashed line in left). Bathymetric data compiled by William Chadwick, Oregon State University
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1
(left) Map of the Galápagos region with 1-km resolution, model grid lines (thin gray lines), and grid points used to represent the Galápagos Islands in the model (heavy black lines). (right) Vertical section of subsurface topography across Isla Isabela, constructed from multiple sources of bathymetric measurements along 91.25°W (black dashed line in left). Bathymetric data compiled by William Chadwick, Oregon State University
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Time series of SST anomaly (°C) in the Niño-4 (red), Niño-3 (black), and Niño-1 + 2 (blue) regions for idealized forced experiments NoGF (dashed) and GF (solid).
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Time series of SST anomaly (°C) in the Niño-4 (red), Niño-3 (black), and Niño-1 + 2 (blue) regions for idealized forced experiments NoGF (dashed) and GF (solid).
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1
Time series of SST anomaly (°C) in the Niño-4 (red), Niño-3 (black), and Niño-1 + 2 (blue) regions for idealized forced experiments NoGF (dashed) and GF (solid).
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Time series of mixed layer heat budget anomalies (W m−2): net heat flux (NHF′), entrainment mixing (EMX′), zonal thermal advection (ZA′), meridional advection (MA′), and latent heat flux (LAT′) in the Niño-3 region for experiments (left) NoGF and (right) GF. Longwave and sensible heat flux anomalies are relatively small (±5 W m−2) and thus omitted from the figure.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Time series of mixed layer heat budget anomalies (W m−2): net heat flux (NHF′), entrainment mixing (EMX′), zonal thermal advection (ZA′), meridional advection (MA′), and latent heat flux (LAT′) in the Niño-3 region for experiments (left) NoGF and (right) GF. Longwave and sensible heat flux anomalies are relatively small (±5 W m−2) and thus omitted from the figure.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1
Time series of mixed layer heat budget anomalies (W m−2): net heat flux (NHF′), entrainment mixing (EMX′), zonal thermal advection (ZA′), meridional advection (MA′), and latent heat flux (LAT′) in the Niño-3 region for experiments (left) NoGF and (right) GF. Longwave and sensible heat flux anomalies are relatively small (±5 W m−2) and thus omitted from the figure.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

(top) Monthly time series of Niño-3 (150°–90°W, 5°S–5°N) SSTA (°C) for 68 years of integration from hybrid coupled experiments NoGC and GC. (bottom) Power spectra [rms as a function of period (years)] for the time series in the top panel plus that of 68 years (1935–2002) of observationally based SST data (Smith and Reynolds 2004).
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

(top) Monthly time series of Niño-3 (150°–90°W, 5°S–5°N) SSTA (°C) for 68 years of integration from hybrid coupled experiments NoGC and GC. (bottom) Power spectra [rms as a function of period (years)] for the time series in the top panel plus that of 68 years (1935–2002) of observationally based SST data (Smith and Reynolds 2004).
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1
(top) Monthly time series of Niño-3 (150°–90°W, 5°S–5°N) SSTA (°C) for 68 years of integration from hybrid coupled experiments NoGC and GC. (bottom) Power spectra [rms as a function of period (years)] for the time series in the top panel plus that of 68 years (1935–2002) of observationally based SST data (Smith and Reynolds 2004).
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

First two (three) EOF patterns of SSTA for hybrid coupled experiments NoGC (GC), computed over the full 68 years of integration in each experiment. Also shown are the (bottom left) corresponding principal components and Niño-3 SSTA indices for each experiment. Time indices only shown for the final 27 years of integration for clarity.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

First two (three) EOF patterns of SSTA for hybrid coupled experiments NoGC (GC), computed over the full 68 years of integration in each experiment. Also shown are the (bottom left) corresponding principal components and Niño-3 SSTA indices for each experiment. Time indices only shown for the final 27 years of integration for clarity.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1
First two (three) EOF patterns of SSTA for hybrid coupled experiments NoGC (GC), computed over the full 68 years of integration in each experiment. Also shown are the (bottom left) corresponding principal components and Niño-3 SSTA indices for each experiment. Time indices only shown for the final 27 years of integration for clarity.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Leading mode of sea level anomaly for hybrid coupled experiments (left) NoGC and (right) GC, computed over the full 68 years of integration in each experiment. Also shown are the (bottom) corresponding principal components and Niño-3 SSTA indices for each experiment. Time indices only shown for the final 27 years of integration for clarity.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Leading mode of sea level anomaly for hybrid coupled experiments (left) NoGC and (right) GC, computed over the full 68 years of integration in each experiment. Also shown are the (bottom) corresponding principal components and Niño-3 SSTA indices for each experiment. Time indices only shown for the final 27 years of integration for clarity.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1
Leading mode of sea level anomaly for hybrid coupled experiments (left) NoGC and (right) GC, computed over the full 68 years of integration in each experiment. Also shown are the (bottom) corresponding principal components and Niño-3 SSTA indices for each experiment. Time indices only shown for the final 27 years of integration for clarity.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Composite evolution of Niño-3 SSTA (°C), SLA PC, SST PC1 and PC2, time rate of change of Niño-3 mixed layer heat content (109 J m−2 s−1), and zonal wind stress anomaly (dyn cm−2) averaged between 5°S and 5°N;) for the last 27 years of hybrid coupled experiments (a) NoGC and (b) GC. Also shown is a simultaneous comparison of the Niño-3 SSTA composites for the (c) two experiments and (d) that for the time rate of change of the composite Niño-3 SSTA, including that computed from observations (Reynolds et al. 2002). Composite time zero refers to the month in which the maximum Niño-3 SSTA was simulated or observed.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Composite evolution of Niño-3 SSTA (°C), SLA PC, SST PC1 and PC2, time rate of change of Niño-3 mixed layer heat content (109 J m−2 s−1), and zonal wind stress anomaly (dyn cm−2) averaged between 5°S and 5°N;) for the last 27 years of hybrid coupled experiments (a) NoGC and (b) GC. Also shown is a simultaneous comparison of the Niño-3 SSTA composites for the (c) two experiments and (d) that for the time rate of change of the composite Niño-3 SSTA, including that computed from observations (Reynolds et al. 2002). Composite time zero refers to the month in which the maximum Niño-3 SSTA was simulated or observed.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1
Composite evolution of Niño-3 SSTA (°C), SLA PC, SST PC1 and PC2, time rate of change of Niño-3 mixed layer heat content (109 J m−2 s−1), and zonal wind stress anomaly (dyn cm−2) averaged between 5°S and 5°N;) for the last 27 years of hybrid coupled experiments (a) NoGC and (b) GC. Also shown is a simultaneous comparison of the Niño-3 SSTA composites for the (c) two experiments and (d) that for the time rate of change of the composite Niño-3 SSTA, including that computed from observations (Reynolds et al. 2002). Composite time zero refers to the month in which the maximum Niño-3 SSTA was simulated or observed.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Solutions to Eqs. (3) and (4) showing the dependence of Kelvin wave phase speed cKe (dashed) and Kelvin wave zonal geostrophic current anomaly uKe (bold) on the depth H of the upper layer.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Solutions to Eqs. (3) and (4) showing the dependence of Kelvin wave phase speed cKe (dashed) and Kelvin wave zonal geostrophic current anomaly uKe (bold) on the depth H of the upper layer.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1
Solutions to Eqs. (3) and (4) showing the dependence of Kelvin wave phase speed cKe (dashed) and Kelvin wave zonal geostrophic current anomaly uKe (bold) on the depth H of the upper layer.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Equatorial time–longitude plot of the projection of Kelvin waves onto zonal geostrophic current (contoured every 0.2 m s−1) in the idealized forced experiment NoGF and the difference between cases, GF minus NoGF.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Equatorial time–longitude plot of the projection of Kelvin waves onto zonal geostrophic current (contoured every 0.2 m s−1) in the idealized forced experiment NoGF and the difference between cases, GF minus NoGF.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1
Equatorial time–longitude plot of the projection of Kelvin waves onto zonal geostrophic current (contoured every 0.2 m s−1) in the idealized forced experiment NoGF and the difference between cases, GF minus NoGF.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Ratio of the standard deviaition of the Kelvin wave zonal geostrophic velocity anomaly signal as a function of longitude for experiments (top) GC to NoGC and (bottom) the difference in the mean state of the equatorial Pacific with and without the Galápagos Islands: annual mean difference in temperature at depth (shaded) and the annual mean distribution of isotherms in the NoGC case (contour interval 0.5°C). The heavy black line is the 20°C isotherm.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Ratio of the standard deviaition of the Kelvin wave zonal geostrophic velocity anomaly signal as a function of longitude for experiments (top) GC to NoGC and (bottom) the difference in the mean state of the equatorial Pacific with and without the Galápagos Islands: annual mean difference in temperature at depth (shaded) and the annual mean distribution of isotherms in the NoGC case (contour interval 0.5°C). The heavy black line is the 20°C isotherm.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1
Ratio of the standard deviaition of the Kelvin wave zonal geostrophic velocity anomaly signal as a function of longitude for experiments (top) GC to NoGC and (bottom) the difference in the mean state of the equatorial Pacific with and without the Galápagos Islands: annual mean difference in temperature at depth (shaded) and the annual mean distribution of isotherms in the NoGC case (contour interval 0.5°C). The heavy black line is the 20°C isotherm.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Solutions to Eq. (5) initialized for values of T ′east of 0.5°, 1.0°, 1.5°, and 2.0°C.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Solutions to Eq. (5) initialized for values of T ′east of 0.5°, 1.0°, 1.5°, and 2.0°C.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1
Solutions to Eq. (5) initialized for values of T ′east of 0.5°, 1.0°, 1.5°, and 2.0°C.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Composite evolution of upper pycnocline mass convergence (normalized units), Niño-3 SSTA (°C), and the time rate of change of Niño-3 mixed layer heat content (109 J m−2 s−1) for the last 27 years of hybrid coupled experiments (a) NoGC and (b) GC. Composite time zero refers to the month of the maximum Niño-3 SSTA.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Composite evolution of upper pycnocline mass convergence (normalized units), Niño-3 SSTA (°C), and the time rate of change of Niño-3 mixed layer heat content (109 J m−2 s−1) for the last 27 years of hybrid coupled experiments (a) NoGC and (b) GC. Composite time zero refers to the month of the maximum Niño-3 SSTA.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1
Composite evolution of upper pycnocline mass convergence (normalized units), Niño-3 SSTA (°C), and the time rate of change of Niño-3 mixed layer heat content (109 J m−2 s−1) for the last 27 years of hybrid coupled experiments (a) NoGC and (b) GC. Composite time zero refers to the month of the maximum Niño-3 SSTA.
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Phase diagram (scatterplot with consecutive points connected by lines) defined by Niño-3 SSTA (°C) along the y axis and zonal wind stress anomaly (dyn cm−2) averaged between 5°S and 5°N) along the x axis for 27 years of hybrid coupled experiments (a) NoGC, (b) GC, and (c) 25 years (1982–2006) of observations (Reynolds et al. 2002; Kalnay et al. 1996).
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1

Phase diagram (scatterplot with consecutive points connected by lines) defined by Niño-3 SSTA (°C) along the y axis and zonal wind stress anomaly (dyn cm−2) averaged between 5°S and 5°N) along the x axis for 27 years of hybrid coupled experiments (a) NoGC, (b) GC, and (c) 25 years (1982–2006) of observations (Reynolds et al. 2002; Kalnay et al. 1996).
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1
Phase diagram (scatterplot with consecutive points connected by lines) defined by Niño-3 SSTA (°C) along the y axis and zonal wind stress anomaly (dyn cm−2) averaged between 5°S and 5°N) along the x axis for 27 years of hybrid coupled experiments (a) NoGC, (b) GC, and (c) 25 years (1982–2006) of observations (Reynolds et al. 2002; Kalnay et al. 1996).
Citation: Journal of Physical Oceanography 38, 11; 10.1175/2008JPO3848.1