1. Introduction
El Niño–Southern Oscillation (ENSO) events are characterized by prominent changes to the tropical climate system. Under ENSO warm phase (or El Niño) conditions, widespread warming of both the troposphere and surface is evident across the Pacific as well as throughout the rest of the tropical latitudes (hereafter the “remote Tropics”) and is accompanied by a spatially complex distribution of precipitation anomalies. In the Pacific positive precipitation anomalies occur in the vicinity of the equator from the date line eastward to the coastline of South America and are attributed to the presence of positive sea surface temperature anomalies (SSTAs) in the central and eastern Pacific (Quinn et al. 1978; Ropelewski and Halpert 1987). Throughout the remote Tropics, on the other hand, widespread negative precipitation anomalies are characteristic during El Niño. Both observational and modeling studies have noted the occurrence of regional precipitation deficits during El Niño in the Amazon basin (Zeng and Neelin 1999), northeastern Brazil (Hastenrath and Heller 1977), the tropical Atlantic (Giannini et al. 2001), the Sahel region of West Africa (Janicot et al. 2001), and the South Asian monsoon region (Webster and Yang 1992; Torrence and Webster 1999).
Although the features of the ENSO tropical teleconnection have been widely documented, the mechanisms underlying the remote ENSO teleconnection, especially the precipitation response, are not completely understood. The leading paradigm for El Niño–forced precipitation changes over the tropical belt is the “anomalous Walker circulation,” the zonal circulation pattern that results from zonal displacements in convection over the tropical Pacific (Kidson 1975; Saravanan and Chang 2000). The conceptual framework of the anomalous Walker circulation is often attributed to Gill (1980), who argued for a simplification of the vertical structure of the convectively driven circulation in the Tropics. Assuming a single baroclinic mode, Gill reduced the tropical dynamics to a system of linearized, damped shallow-water equations on an equatorial beta plane. Gill derived analytical solutions for the circulations driven by anomalous imposed diabatic forcing, including the structure of forced Kelvin and Rossby waves and the associated subsidence field away from the forcing region. In the context of the ENSO tropical teleconnection, the reduced precipitation throughout the remote Tropics during El Niño is attributed to the increase of subsidence forced by anomalous uplift over the warm Pacific SSTA (Wu and Newell 1998; Klein et al. 1999; Su et al. 2001; Su and Neelin 2002).
While conceptually powerful, the Gill model has significant shortcomings when applied to the tropical ENSO teleconnection. Most significant is its neglect of moist thermodynamics, radiation, and turbulent fluxes that ultimately determine the climate (including precipitation) response to El Niño. As the Gill framework is essentially a dynamical construct, it only predicts precipitation changes to the extent that these may be inferred from changes to subsidence. A characteristic feature of the Gill solution to an idealized equatorially symmetric heating source is the relatively spatially smooth nature of the subsidence field away from the forcing region (cf. Fig. 1d of Gill 1980), implying that subsidence is relatively uniformly distributed over the remote Tropics. However, a cursory examination of the anomalous tropical pressure velocity field during either El Niño or La Niña conditions (e.g., see Su and Neelin 2002) shows that the anomalous subsidence field exhibits a nontrivial spatial structure (Fig. 1). During El Niño events (Fig. 1b), negative (ascent) anomalies prevail over the central and eastern Pacific. Outside of this region, there is a tendency for positive (descent) anomalies, but the structure is rather heterogeneous. The distribution of anomalous subsidence appears to be linked to the mean convective conditions of the remote tropical region and, in fact, points to the important role of moist thermodynamic feedbacks. Indeed, recent studies (e.g., Su et al. 2001; Chiang and Sobel 2002, hereafter CS02; Su and Neelin 2002; Neelin et al. 2003; Su et al. 2004) have recognized the significant role of moist thermodynamics and have begun to approach the tropical ENSO teleconnection with a moist thermodynamic emphasis.
The motivation and approach for our current study derives from CS02, who sought to highlight the role of moist thermodynamics by simplifying the horizontal dynamics of the teleconnection problem. CS02 assumed that the effect of horizontal tropical dynamics is to spread the tropospheric temperature warming caused by El Niño uniformly across the tropical latitudes, invoking the “weak temperature gradient” (WTG) approximation (e.g., see Sobel and Bretherton 2000). Since the Coriolis effect is small near the equator, the tropical troposphere is unable to maintain significant horizontal temperature gradients. CS02 proceeded by imposing tropospheric temperature anomalies within a single-column model (SCM) equipped with parameterizations of moist convection, radiation, and boundary layer physics and coupled to a passive thermal ocean. With the tropospheric temperature profile imposed, the climate response (including subsidence) of the remote Tropics is determined simply by thermodynamic energy balance constraints. Among the notable results of CS02 were a relatively simple explanation for the remote tropical surface warming during El Niño (further explored in Chiang and Lintner 2005) and a proposed mechanism for remote tropical precipitation anomalies based on the disequilibrium between the free tropospheric and boundary layer moist static energies.
We extend the work of CS02 by investigating the ENSO response of the entire tropical atmosphere using a set of SCMs connected through the WTG requirement. Sobel and Bretherton (2000) showed that an intermediate level complexity model [the Quasi-equilibrium Tropical Circulation Model (QTCM); see section 2] modified in this manner was able to reproduce a reasonable tropical climatology; we use the same model and a similar procedure to model perturbations to tropical climate during El Niño. Since the WTG approximation by itself constrains only gradients in tropical tropospheric temperature, we determine the absolute level of tropospheric temperature through a mass balance requirement that the net divergence anomaly summed over the tropical belt is zero; similar approaches to “close” the problem have previously been used in more idealized models employing WTG, for example, Shaevitz and Sobel (2004). The mass balance constraint follows from the assumption that Pacific region ascent (descent) anomalies associated with El Niño (La Niña) are mostly compensated for by descent (ascent) elsewhere in the Tropics, an aspect of the anomalous Walker circulation that can be justified on the basis of the tropical waveguide properties of equatorial atmospheric dynamics (Gill 1980). An examination of the observed divergence anomalies associated with ENSO supports this view (see section 3). In what follows, we will refer to the constraints on the anomalous tropical tropospheric temperature and divergence fields collectively as the “WTG framework.”
We will demonstrate that the gross features of the teleconnection—for example, the tropospheric temperature and precipitation response averaged over the remote tropical region—appear to be well simulated within the WTG framework, as are the spatial details of the precipitation response in the tropical Pacific. By contrast, the spatial details of the remote tropical precipitation response are less adequately simulated; we suggest that this disagreement may result from the neglect of features such as horizontal temperature gradients. The organization of the remainder of this paper is as follows. After introducing the models and methodology and providing some justification for the WTG approach in sections 2 and 3, we briefly document the principal features of the mean precipitation field simulated by the standard and WTG versions of the QTCM (section 4). An in-depth discussion of the teleconnected responses to El Niño (and La Niña) forcing conditions in both model versions follows in section 5. We then present results of perturbative simulations with increased atmospheric greenhouse gas loading (section 6); the purpose of these simulations is to demonstrate further the applicability of the WTG approach as well as to highlight the distinct influences of temperature and large-scale subsidence changes during El Niño. In section 7, we briefly explore the impact of anomalous temperature gradients on the characteristics of the response to El Niño conditions. We conclude with a summary of our principal findings and a discussion of the implications of our study for the understanding of ENSO tropical teleconnectivity (section 8).
2. Models and methodology
a. The Quasi-equilibrium Tropical Circulation Model
The QTCM is an intermediate level complexity model of the tropical atmosphere developed by the Climate Systems Interactions group at the University of California, Los Angeles (Neelin and Zeng 2000; Zeng et al. 2000). The QTCM incorporates all relevant dynamics and physical parameterizations that are necessary to determine the tropical climate state, that is, atmospheric dynamics, convection, clouds, radiation, and turbulent fluxes. Our use of QTCM is motivated by the ease with which it can be modified to employ the WTG framework (as we show below); moreover, the QTCM contains all of the important physics (including moist thermodynamics) that we assume to be of importance for the teleconnection problem. The QTCM has been used previously to investigate many aspects of tropical climate, including monsoons (Chou et al. 2001) and tropical ENSO teleconnections (Su et al. 2001; Su and Neelin 2002).








b. The WTG version of QTCM




Our approach treats the temperature field as two distinct components, namely an imposed, heterogeneous mean field and an adjustable, homogeneous offset. As such, our application of WTG only applies to perturbations about the tropospheric temperature climatology and not to the climatology itself—this is not unreasonable since our interest is in perturbations of the tropical climate to ENSO. It should be noted that we implemented WTG fully by requiring a spatially homogeneous mean tropospheric temperature field; while the full WTG approach did not perform as well in simulating the mean climate, qualitatively similar anomalous climate behavior was obtained with the full WTG approach. We also point out that, strictly speaking, the application of WTG does not require temperature gradients to vanish. Rather, the approximation involves imposing the limit of zero-valued temperature gradients in the thermodynamic equation [Eqs. (1) and (2)] only. Temperature gradients also appear in the horizontal momentum equations, but WTG places no constraints on temperature gradients as they influence momentum. However, our application of WTG effectively assumes a zero-valued anomalous horizontal circulation field, so the issue of nonzero temperature gradients in the momentum equations is neglected. In simpler applications of WTG (such as the two-column model of Shaevitz and Sobel 2004), it is, in fact, possible to infer a posteriori a circulation consistent with WTG by invoking continuity; the implementation of a WTG-consistent circulation for a model setup with N > 2 columns is less straightforward.
c. Overview of simulation methodology
Simulations of both ENSO and greenhouse gas (GHG) forcing scenarios were performed with each version of the QTCM (hereafter, we refer to the QTCM configured with the WTG assumptions as simply the “WTG version” and the non-WTG configuration as the “standard version”). The horizontal resolution used for all simulations was 5.625° × 3.75°, with the meridional domain spanning 76.875°S–76.875°N. For the ENSO forcing experiments, SSTs were imposed throughout the Pacific from 110°E to the western coastline of the Americas and from 30°S to 30°N. At all other oceanic grid points, the atmospheric model was coupled to a passive slab ocean model with a spatially invariant mixed-layer depth (MLD) of 50 m. Additionally, a Q-flux correction was applied at each slab ocean grid point to maintain a realistic SST climatology (Chiang et al. 2003). Both control simulations, with only the SST climatology imposed, and perturbation simulations, with SST climatology plus anomalies imposed, were performed. (Anomalies of model fields are defined as differences between the perturbation and control simulations.) For the greenhouse gas forcing experiments, the QTCM was coupled to a Q-flux-corrected slab ocean model at all oceanic grid points.
To simplify the interpretation of the results, only perpetual January conditions were considered; that is, seasonality was suppressed in the climatological SST, surface albedo, and solar radiation forcing fields, and mean January conditions were assumed. In all forcing scenarios, ensembles consisting of 10 members were obtained; each ensemble member was initialized from a set of unique, self-consistent atmospheric and land surface initial conditions obtained from a lengthy control integration. Individual ensemble member integrations of 8 years were simulated with the output fields archived as monthly means.
3. Justification for the WTG approach
Justification for the weak temperature gradient constraint has been thoroughly argued in previous studies [e.g., see Sobel and Bretherton (2000) and references therein for the tropical mean state, and CS02 for the ENSO problem], so we do not go into the details here. To assess the validity of the mass balance constraint—that perturbations in ascent induced by El Niño are essentially entirely compensated for by subsidence within the tropical band—we present composite analysis results for the annual-mean 500-mb pressure velocity field, obtained from the National Centers for Environmental Prediction–National Center for Atmospheric Research (NCEP–NCAR) reanalysis dataset and averaged over selected domains in the Tropics for the period 1950–2002 (Table 1).
Consider the vertical velocity field averaged over 20°S–20°N. In both the Pacific and remote regions, the areal-mean pressure velocity is negative, corresponding to ascending motion in the rising branch of the Hadley circulation. During El Niño phases, there is an ∼5% increase in ascent (from −0.96 × 10−2 Pa s−1 to −1.01 × 10−2 Pa s−1) in the tropical Pacific that is compensated almost entirely by an increase in the subsidence (i.e., the magnitude of the remote-averaged pressure velocity decreases from 0.37 × 10−2 Pa s−1 to 0.30 × 10−2 Pa s−1) over the remote Tropics so that the mean ascent over the entire Tropics is relatively unchanged. A similar picture, but with Pacific and remote regional anomalies of opposing sign, emerges for La Niña conditions. Thus, we conclude that the vertical velocity field, when averaged over the entire Tropics (at least for a reasonable range of latitudes), is relatively insensitive to ENSO phase. Note that, for an averaging interval of 30°S–30°N, the average of the vertical velocity computed over all longitudes is much closer to zero, reflecting the cancellation of the ascending and descending branches of the Hadley circulation in the meridional direction.
In a related vein, we note the pronounced tendency of observed remote tropical precipitation anomalies to anticorrelate with the equatorial Pacific anomalies. Over the period 1979–99, for example, time series of the aggregate, monthly mean Pacific and remote tropical precipitation anomalies from the Climate Prediction Center (CPC) Merged Analysis of Precipitation (CMAP) dataset are anticorrelated beyond the 99.9% level (r = −0.30). Given the association between ascent (subsidence) and enhanced (reduced) precipitation, an anomalous zonal dipole behavior in precipitation is not surprising. Similar regionally antiphased rainfall anomalies are also evident in models. For example, a multidecadal (1950–2000) integration of the QTCM with imposed Pacific SSTs exhibits strongly pronounced antiphasing between its Pacific and remote regional average precipitation anomalies (r = −0.93); other models [including, for example, the National Center for Atmospheric Research (NCAR) Community Climate Model version 3.10 (CCM3; see Kiehl et al. (1998)] produce similar behavior, although the degree of anticorrelation is variable.
4. Simulated precipitation climatologies
We first briefly document the principal features of the precipitation climatologies for the WTG and standard versions of the QTCM [see Sobel and Bretherton (2000) for a more thorough discussion of the mechanisms underlying the establishment of the mean state precipitation climatology in the single-column version of QTCM]. Global maps of the mean precipitation fields for the control runs of both the standard and WTG QTCM frameworks, averaged over the period when the simulations have equilibrated (months 61–96), are presented in Figs. 2a and 2b, respectively. Given the simplifications inherent in the WTG version of QTCM, the distribution and magnitudes of the WTG precipitation climatology match the standard QTCM climatology remarkably well. Focusing on the tropical regions, both versions exhibit centers of deep convection over the Amazon basin, Africa, and the equatorial Indian Ocean/Maritime Continent/western Pacific regions. For the continental convection zones, peak monthly mean rainfall rates of up to 14 mm day−1 are evident, with rainfall rates reaching upward of 20 mm day−1 near the Maritime Continent.
There are however some differences between the simulated climatologies: in particular, the WTG QTCM simulates slightly higher precipitation rates (typically 2 mm day−1) in the regions of strongest mean convection and lower rates on the margins of the strongest convection zones. In other words, the climatological convection field in the WTG QTCM tends to be spatially more sharply defined than in the standard version. Such differences likely arise from the lack of interactive high frequency coupling between thermodynamics and dynamics (i.e., transient humidity and temperature fluxes) that would tend to “smear” the edges of deep convective regions. Despite these differences, the overall correspondence between the climatologies should permit reasonable comparisons of the anomalous climate response to ENSO in the two QTCM versions.
5. Standard and WTG QTCM simulations of ENSO forcing
To produce the El Niño forcing response in both the standard and WTG versions of the QTCM, the SSTA field corresponding to the observed anomalies for January 1998—the peak month of the 1997/98 El Niño (the “El Niño of the century”)—was added to January SST climatology in the Pacific forcing region. (Note that the SSTAs were linearly suppressed to zero between 30°–20°S and 20°–30°N, 110°–125°E.) For the WTG QTCM, the diabatic vertical velocity field and uniform temperature offsets were computed via Eqs. (5) and (6), respectively. In these equations, a meridional averaging interval of 30°S–30°N was employed; the use of different averaging intervals (e.g., 20°S–20°N) was found to yield qualitatively similar results. As an additional test of the applicability of the WTG framework, we also performed simulations using La Niña phase SSTA. To facilitate comparison to the El Niño scenario, we used as the imposed perturbation forcing the SSTA field from the El Niño simulation but with its polarity reversed. Actual La Niña events, such as the 1989 event, were simulated with qualitatively similar behavior obtained.
a. Gross responses over the remote Tropics
The widespread warming (cooling) of the tropical troposphere and the remote surface and the reduction (increase) in remote precipitation are two of the most recognizable and robust features of the remote tropical response to El Niño (La Niña). Table 2 summarizes the (vertically averaged) tropospheric temperature (TT), surface temperature (Ts), (vertically averaged) tropospheric specific humidity (q), and precipitation (PPT) changes in response to the imposed anomalous Pacific region SST forcing for each QTCM version. Each quantity tabulated has been averaged over the final 36 months of the simulation and represents an areal average over the remote tropical domain spanning 20°S–20°N, 70°W–110°E.
Overall, the agreement of the quantities listed in Table 2 between the standard and WTG QTCM simulations of the El Niño and La Niña forcing scenarios is impressive considering the simplifications associated with WTG: in particular, both the signs and magnitudes of the anomalies in the WTG and standard simulations are similar. For the remote TT response to El Niño conditions, for instance, both QTCM versions simulate tropospheric temperature warming of around ∼0.7 K. Along with the tropospheric temperature increase, surface temperature increases by 0.45 K and specific humidity by 0.28 K. The magnitude of WTG-simulated tropospheric temperature, surface temperature, and specific humidity anomalies all exceed the standard QTCM values by about 11%–16%. The largest discrepancy between the standard and WTG versions, however, is with the remote tropical precipitation deficit, with the WTG precipitation anomaly only 60% as large as the standard QTCM anomaly.
The El Niño phase warming of the remote tropical surface and the increase in remote tropospheric humidity may be understood in the context of the thermodynamic mediation of the remote ENSO teleconnection (CS02; Chiang and Lintner 2005). The linkage between remote surface temperatures and the free tropospheric temperature anomaly is envisioned to occur via moist convective processes: in particular, the covariation of free tropospheric temperature and boundary layer moist static energy (under quasi-equilibrium constraints; see Brown and Bretherton 1997) requires that both surface temperature and humidity increase. Additional nonthermodynamic controls may also exist: for example, changes to cloud type fraction and distribution may impact the surface shortwave and longwave radiative balance. Although the anomalous cloud radiative forcings in the shortwave and longwave largely cancel, the cancellation is not exact, especially regionally. Furthermore, surface wind speed anomalies may alter surface turbulent fluxes. Chiang and Lintner (2005) present the results of some idealized simulations that suggest that these effects, while perhaps regionally significant, are of nonleading-order importance to the gross-scale responses identified here.
The characteristics of the comparison between the standard and WTG La Niña simulation are similar to the El Niño response (but with the signs of the anomalies reversed), except that in the La Niña case the magnitudes of the WTG responses are underestimated relative to the standard simulation. An encouraging aspect of the WTG simulation is that, like the standard simulation, it captures the nonlinearity in the atmospheric response to ENSO phase: while the La Niña SST forcing is exactly reversed from the El Niño forcing, the magnitudes of the La Niña phase climate responses are generally only 30%–40% as large as the El Niño phase responses.
What accounts for the differential response of the two QTCM versions? At least some of the difference can be attributed to the WTG constraint on the anomalous tropospheric temperature perturbation: the warming that originates in the Pacific is distributed isotropically across the entire tropical belt such that the remote tropical region warming in the WTG framework is exaggerated. In some sense, then, a more valid comparison is between the zonal-mean tropospheric temperature anomalies of the standard and WTG QTCM versions rather than the remote regional averages. The zonal-mean TT anomaly in the standard simulation is 0.69 K, showing that the tropospheric temperature difference between the WTG and standard simulations, while still nonzero, is somewhat reduced. Another source of divergence is differences in the spatial distribution of climate anomalies in the remote Tropics, an aspect of the teleconnection that is elaborated in the next subsection.
b. Spatial characteristics of the precipitation and surface temperature response
We now examine the spatial structure of the anomalous precipitation responses to ENSO conditions in each QTCM version. Figure 3 illustrates latitude–longitude maps of the El Niño precipitation anomalies in the standard and WTG frameworks averaged over the final 36 months of each simulation. We present separately the tropical Pacific “forcing” and remote tropical “response” regions because the magnitudes of the tropical Pacific precipitation anomalies are an order of magnitude larger than the remote tropical anomalies. Analogous maps for La Niña phase conditions appear in Fig. 4.
For the Pacific region El Niño anomalies, the spatial characteristics of anomalous precipitation in the standard and WTG frameworks (Figs. 3a and 3c, respectively) are, to lowest order, quite similar. Both model versions simulate enhanced precipitation in the central and eastern portions of the Pacific from the date line eastward to the coast of South America. To the west of this region precipitation is suppressed, as it is in the bounding margins to the north and south. Such patterns of precipitation anomalies match the observations as well as the El Niño phase precipitation anomalies simulated by other models (not shown). The magnitudes of the WTG and standard responses differ significantly, however: in particular, the WTG version simulates a greater maximum precipitation increase (by about a factor of 2) in the anomalous, near-equatorial wet zone (Fig. 3e). Additionally, the positive anomalies in the WTG simulation extend over a greater meridional interval, and the gradient between the maximum anomalies at the center of the wet zone and those on the margins is sharper. Commensurate with the larger positive anomalies in the central and eastern portions of the basin in the WTG framework, the negative anomalies in the western portion of the basin are also more pronounced. However, for the north- and south-flanking negative anomalies, the differences between the QTCM versions are mixed: the deficit to the south is somewhat more extensive in the WTG version with an enhanced peak reduction, while the deficit to the north is far less spatially extensive and weaker in magnitude.
The different Pacific region precipitation responses may be interpreted in terms of the different tropospheric temperature field structure of the two model versions (Fig. 3g). Across the central and eastern (western) Pacific the standard simulation exhibits a larger (smaller) increase in tropospheric temperature than the WTG simulation, which gives rise to an east–west temperature gradient (e.g., the difference between spatial averages of tropospheric temperature for the eastern and western portions of the Pacific basin in the standard version is ∼1 K) that is comparable in magnitude to Pacific region mean temperature increase (the average temperature change over the entire Pacific region is 0.72 K). Because the WTG QTCM imposes temperature anomalies uniformly, it effectively underestimates the El Niño phase temperature rise in the eastern Pacific and overestimates it to the west, the net result of which is a relative amplification of the zonal precipitation anomaly dipole. The poor agreement of Pacific regional precipitation anomaly magnitudes points to the limited validity of the WTG framework in the presence of sizeable temperature gradients.
For the remote tropical region, the standard QTCM simulation of El Niño conditions (Fig. 3b) exhibits a spatially complex precipitation response with localized regions of higher rainfall interspersed among geographically extensive rainfall deficits. As noted in Table 2, the prevailing precipitation response for the remote Tropics as a whole is a reduction in rainfall (see Table 2). Several prominent features of the anomalous precipitation response, including the “bull’s-eye” of negative precipitation anomalies over South America as well as the drying across large portions of the Indian Ocean basin, are consistent with observed El Niño phase behavior (not shown). As in the standard simulation, the WTG QTCM produces a gross-scale deficit in remote tropical precipitation during El Niño, albeit underestimated. However, the spatial characteristics of the WTG remote precipitation response (Fig. 3d) compare relatively poorly with those of the standard simulation (Fig. 3b). In particular, the negative anomalies over the interior northern regions of South America in the standard version are supplanted by positive anomalies in the WTG version; while negative anomalies are present, they are essentially confined to the coastal margin. The positive precipitation anomalies over Africa are stronger and more widespread in the WTG version, and a tongue of positive precipitation anomalies is coincident with the highest rainfall rates over the Indian Ocean. We revisit these differences and assess their implications in sections 6 and 7.
Our conclusions regarding the comparison between the WTG and standard El Niño precipitation responses also hold for the La Niña simulations (Fig. 4): the spatial distribution of the precipitation anomalies is qualitatively similar between the WTG and standard simulations in the tropical Pacific forcing region, but the spatial agreement is poor in the remote Tropics. One notable feature inherent in the WTG simulation—and noted in the standard version as well—is the presence of an asymmetry in the precipitation responses between El Niño and La Niña phases. Geographically, the El Niño and La Niña precipitation anomalies are not strictly anitsymmetric, for example, the large response over the date line in the equatorial Pacific during El Niño is essentially absent in the La Niña simulation. Asymmetry in the atmospheric response to ENSO phase thus appears to be inherent to the physics of the WTG approach. More specifically, the asymmetry does not arise from the nondivergent component of the circulation response, since our application of WTG does not model that circulation response, nor does it develop in response to turbulent flux anomalies associated with ENSO-induced modulation of the surface circulation.
Relatively poor agreement is also evident in the spatial distributions of the WTG and standard El Niño remote tropical surface temperature anomalies (Fig. 5). While there is a net tendency toward warming in the remote Tropics in both the WTG and standard simulations (see Table 2 and section 5a), the warming in the latter is generally concentrated over specific locations, namely the tropical continents and the equatorial Indian Ocean, while the warming in the former is more diffuse. The largest differences in WTG and standard QTCM-simulated surface temperature anomalies (Fig. 5c) are often spatially collocated with precipitation anomaly differences (Fig. 3f), with regions of negative precipitation anomaly differences associated with positive surface temperature anomaly differences. These associations are especially evident over equatorial South America where the WTG version significantly underestimates precipitation compared to the standard simulation and also over the tropical Indian Ocean where the WTG simulation shows increased (decreased) near- (off-)equatorial precipitation. (The standard QTCM exhibits decreased precipitation throughout the tropical Indian Ocean.) The linkage between precipitation and surface temperature may reflect cloud–radiative effects: decreased rainfall produces fewer clouds, allowing increased shortwave heating of the surface.
Other influences neglected in the WTG framework may contribute to the different remote tropical surface temperature responses between the WTG and standard simulations. For instance, since the horizontal circulation is fixed to climatology, changes in the turbulent fluxes associated with surface circulation changes (the wind speed effect on latent and sensible heat fluxes) are not modeled. Another potential source of discrepancy is the lack of a spatially varying anomalous tropospheric temperature response. Since the degree of surface warming appears to be tied to the level of free tropospheric temperature change (according to the mechanism of CS02), differences in the spatial structure of the tropospheric warming between the standard and WTG QTCM simulations may contribute to the differing surface temperature responses. Indeed, inspection of the standard and WTG tropospheric temperature anomaly difference field (Fig. 3h) reveals that the standard QTCM simulates a larger increase in tropospheric temperature over equatorial South America compared to the WTG simulation; the differential tropospheric warming may account for the pronounced anomalous surface temperature difference between the WTG and standard simulations there.
6. Understanding the impacts of tropospheric temperature and subsidence changes during El Niño
We now try to understand the origins of the WTG El Niño response in the remote Tropics by attempting to separate the relative influences of the tropospheric temperature warming and subsidence on the climate. Although the El Niño–induced remote tropical increases in tropospheric temperature and subsidence are causally related and difficult to separate (as adiabatic compression associated with subsidence gives rise to warming), from a precipitation perspective, the effects of tropospheric temperature and subsidence may impact convection through different pathways—that is, tropospheric temperature through stabilization of the atmospheric column and subsidence through divergence of column moisture [see Eq. (7) below and the related discussion]. Within the WTG framework we can attempt to assess the relative roles of these two influences.
Apart from the potential to diagnose the mechanisms of the ENSO tropical teleconnection, consideration of the “pure” tropospheric temperature influence of ENSO separate from the subsidence effect is also of potential interest for determining the tropical rainfall response to increased tropospheric greenhouse gas loading, as both increasing tropospheric greenhouse gas concentrations and El Niño are associated with Tropics-wide tropospheric temperature increases. In fact, because of the warming effect common to greenhouse gases and El Niño, the latter has been suggested as a proxy for the near-surface effects of global warming (Soden 1997). Our basic conclusion here is that the direct effects of temperature and subsidence have significantly different impacts on the remote rainfall response to ENSO, so the value of ENSO as a global warming proxy is questionable.
To investigate the pure tropospheric temperature influence, an increased GHG scenario (denoted “ΔGHG+”) was simulated in each model framework using a 50-m MLD slab ocean model at all oceanic grid points. The CO2 concentration was increased (at the initiation of the perturbation simulation) from its control value of 330 ppmv to a new uniform value everywhere; the level of perturbation CO2 forcing was chosen to yield the same mean equilibrium level of anomalous tropospheric warming as in the WTG El Niño simulation. (The level of perturbation CO2 required to achieve the desired temperature change, 0.78 K, was found to be approximately 1.46 times the standard concentration, or 483 ppmv.) For the WTG ΔGHG+ simulation, the zero divergence anomaly constraint was again enforced over 30°S–30°N.
Anomalous precipitation fields simulated under the ΔGHG+ scenario for the standard and WTG versions of the QTCM are illustrated in Fig. 6. The anomalous rainfall distributions are quite similar between the standard and WTG simulations of the ΔGHG+ scenario (cf. Figs. 6a,b to Figs. 6c,d). In fact, the remote tropical averages of various climate parameters for the two QTCM versions subject to enhanced GHG forcing are found to agree to within a percent (Table 3). The overall agreement between the standard and WTG versions of QTCM may reflect the fact that the forcing induced by increased GHG loading is, to lowest order, spatially uniform, so the dynamical responses tied to pressure gradients (which our WTG framework does not model) are negligible. In terms of the response to the greenhouse gas forcing itself, our results are consistent with those previously reported by Neelin et al. (2003) and Chou and Neelin (2004) using the QTCM. We find that extensive regions of positive precipitation anomalies are observed to occur in the vicinity of the regions of strongest mean convection and drying in marginal convection regions adjacent to the main convective centers, most notably along the eastern coastline of South America, the western coastline of Africa, and across the north and south tropical Indian Ocean.
Neelin et al. (2003) and Chou and Neelin (2004) ascribed the precipitation response to anomalous GHG forcing to the action of two related mechanisms dubbed “upped ante” and “rich-get-richer.” Under the first of these, the increase in tropospheric temperature arising from increased GHG is thought to increase the thermal barrier for convection by stabilizing the troposphere. Deep convective zones, characterized by convergent mean flow, can overcome the upped ante for convection by outcompeting marginal zones for moisture so that the convectively rich become convectively richer. Some elements of these mechanisms (e.g., the increase in precipitation over equatorial Africa) are discernible in the El Niño simulations discussed in section 5.
We now compare the climate responses of the WTG El Niño and ΔGHG+ simulations over the remote tropical region, given that both forcing scenarios exhibit the same uniform tropospheric temperature warming but that they differ in the presence or lack of forced large-scale subsidence. The intriguing result of this comparison is that, while the aggregate remote tropical precipitation anomalies are of opposite sign for the WTG El Niño and ΔGHG+ forcing scenarios, the spatial patterns of the anomalous remote tropical precipitation fields clearly resemble one another, as can be seen by direct comparison of Figs. 6d and 3d. These results suggest that the operation of distinct influences associated with the remote tropical tropospheric temperature and large-scale subsidence increases during El Niño give rise to, respectively, a similar spatial distribution and a dissimilar level of gross-scale anomalous precipitation relative to the ΔGHG+ forcing case. In other words, the presence of a (uniform) temperature increase in the WTG El Niño simulation imparts a GHG-like structure to the spatial distribution of precipitation anomalies over the remote tropical region. However, the presence of a forced large-scale descent anomaly—which is required to balance anomalous ascent in the Pacific—produces the remote-averaged offset in the level of anomalous precipitation.
We can demonstrate the aforementioned aspects of the anomalous precipitation behavior more clearly using a binned averaging procedure to examine changes to the precipitation field as a function of mean-state precipitation. Our motivation for analyzing the data in this way follows from the observation that the sign of the precipitation anomalies in the WTG (or standard) ΔGHG+ scenario appears to depend on the mean-state precipitation, changing from negative (for marginally convective regions) to positive (for deep convective regions). By contrast, for El Niño conditions, the anomalous remote tropical precipitation response appears to be (at least in the standard QTCM) relatively insensitive to the mean precipitation state, that is, negative precipitation anomalies are geographically widespread.
Binned average El Niño (solid lines) and ΔGHG+ (dashed lines) precipitation histograms for the standard and WTG QTCM appear in Figs. 7a and 7b, respectively. Here, the anomalies have been averaged over mean precipitation bins of width 2 mm day−1 and aggregated over the entire remote tropical region (20°S–20°N, 70°W–110°E). The results have been further averaged over the final 36 months of the simulations, with the error bars representing the 2σ level of each bin average. Comparing the results for the standard QTCM (Fig. 7a), it can be seen that the El Niño histogram is negative for nearly all mean precipitation values while the ΔGHG+ histogram is negative below ∼6 mm day−1 and positive above, clearly revealing the distinct nature of the anomalous precipitation responses for the two forcing scenarios.
For the WTG simulations (Fig. 7b), the ΔGHG+ histogram (dashed line) essentially lies on top of the standard ΔGHG+ histogram. By contrast, the standard and WTG El Niño histograms (Fig. 7b) are rather dissimilar, as the latter exhibits a more GHG-like structure: the precipitation anomalies of the WTG El Niño simulation transition from negative to positive with increasing mean-state precipitation but with an offset that depresses the entire histogram relative to the WTG ΔGHG+ result. Although fairly constant for all mean values, the offset is largest at intermediate mean precipitation rates (i.e., 2–6 mm day−1). Again, it is likely that the differences between the standard and WTG QTCM El Niño simulations are attributable to effects neglected by the WTG approach, such as anomalous spatial gradients in temperature.


The tropospheric drying (moistening) effect of subsidence (uplift) acts to reduce (increase) the column humidity through column divergence (convergence) of moisture and, therefore, lowers (raises) the Betts–Miller reference temperature profile. Examination of a longitudinal profile of atmospheric humidity in the WTG QTCM El Niño and ΔGHG+ scenarios (Fig. 8) demonstrates the linkage between vertical motion and humidity. The humidity change of the WTG ΔGHG+ simulation (triangles) is observed to be longitudinally quasi-uniform; the localized regional maxima in the humidity profile correspond to regions of climatological deep convection, where precipitation and vertical motion are increased. On the other hand, the humidity profile for the WTG El Niño simulation (squares) is sharply peaked in the central and eastern Pacific where strong anomalous convection, locally forced by the imposed SSTA, is present. Outside of the tropical Pacific, the humidity profile of the El Niño simulation lies below the level of the ΔGHG+ simulation, suggesting the suppression of specific humidity by mean subsidence over the remote Tropics as a whole in the El Niño simulation. We conclude that an important effect of the anomalous divergence induced by El Niño conditions in the Pacific region is the establishment and maintenance of a lower level of remote tropospheric humidity than that which occurs in the presence of the temperature-only forcing associated with perturbed GHG.


7. Influence of anomalous temperature gradients in the WTG framework
One potential source of the poor representation of the spatial distribution of El Niño–induced precipitation anomalies in the WTG simulations is the assumption of no anomalous temperature gradients. Given that some tropospheric temperature gradients may develop in response to El Niño forcing (see Figs. 4g and 4h), it is of interest to determine what impact the addition of anomalous temperature gradients to the WTG simulation may have on the climate response to ENSO. To do this, we simply added to the WTG El Niño forcing scenario the spatial structure of the anomalous temperature field as simulated by the standard QTCM. Before adding the temperature gradient anomalies, a mean temperature anomaly corresponding to the areal average over the 30°S–30°N strip used to compute the temperature offset was first removed such that the areal-mean of the imposed temperature gradient perturbation was zero.
The ad hoc inclusion of anomalous temperature gradients produces an anomalous precipitation response in the Pacific (Fig. 9a) that more closely matches the standard simulation (Fig. 3a) than does the WTG El Niño simulation lacking anomalous temperature gradients (Fig. 3c). In particular, the zone of positive precipitation anomalies in the central and eastern Pacific is reduced in extent and magnitude with the inclusion of an anomalous temperature gradient field. The near-equatorial negative precipitation anomaly is also reduced in magnitude and extent, while the north-flanking negative anomaly is enhanced. Some improvements are also evident over the remote region (Fig. 9b): in particular, there is a more pronounced penetration of negative anomalies into the interior of South America.
Of course, not all regions show improvements in the precipitation response with the addition of anomalous temperature gradients. Over Africa and the Indian Ocean, the precipitation response in Fig. 9b still essentially resembles the WTG simulation lacking anomalous gradients. However, temperature gradients are relatively weak over this portion of the remote Tropics, suggesting that something other than spatial structure of the temperature field may contribute to the anomalous precipitation behavior. We speculate that circulation effects not modeled in the WTG framework, particularly wind speed changes to the latent heat flux, may impart some of the structure to the standard simulation precipitation field, particularly over the eastern Indian Ocean basin, although an understanding of the details of these effects is beyond the scope of the present study.
8. Conclusions
In this paper, we explored the reorganization of the tropical climate during El Niño using a weak temperature gradient (WTG) framework, with an emphasis on the mechanisms giving rise to the ENSO remote tropical precipitation teleconnection. The fundamental assumptions at the heart of this approach are 1) a spatially homogeneous perturbation temperature change and 2) an anomalous mass conservation constraint applied to the entire tropical belt. In the QTCM, the intermediate-level complexity model used in this study, application of the WTG assumptions simplifies the model by eliminating the prognostic equations for horizontal momentum and replacing the prognostic equation for temperature with a diagnostic equation for divergence and a single anomalous temperature tendency equation closed by the mass balance constraint.
ENSO simulations performed with the QTCM modified according to the WTG framework were found to simulate an appropriate level of tropospheric temperature change as well as a reasonable spatial distribution of precipitation anomalies in a gross-scale sense, demonstrating that the WTG framework can account for the basic features of the tropical response to ENSO forcing conditions. We showed additionally that the WTG QTCM realistically simulates the climate response to enhanced greenhouse gas forcing, at least in comparison to the standard version of the same model. We further used the WTG framework to show how the differential precipitation responses to GHG and El Niño forcing may be tied to fundamental differences in the nature of the forcing, that is, temperature only for GHG and temperature with large-scale-forced subsidence for El Niño.
Despite the agreement of the gross-scale El Niño precipitation deficits in the standard and WTG versions of the QTCM, the regional-scale features of the anomalous precipitation (and surface temperature) field in the remote Tropics manifest some marked contrasts. For example, while the standard QTCM version was found to simulate a widespread reduction of precipitation across the remote region in response to El Niño forcing conditions, the WTG version yielded a more GHG-like pattern of enhanced precipitation over deep convective zones and reduced precipitation along the margins of these zones. The addition of anomalous temperature gradients to the WTG simulation somewhat improved the spatial agreement between the WTG and QTCM precipitation anomaly fields.
Significantly, we suggest that the WTG approach outlined here may provide the underpinnings of a conceptual framework for adjustment of the tropical climate to ENSO, in particular the remote tropical response. The canonical conceptual framework of the ENSO teleconnection, the Gill (1980) model, has proven successful in many ways, especially in terms of understanding the surface and upper-tropospheric atmospheric circulation anomalies during El Niño and the gross changes to the zonal Walker circulation given the anomalous convective heating associated with warm SSTA. However, the Gill model lacks the thermodynamics necessary to understand what the climate response of the remote Tropics is beyond that which can be inferred from the linkages to anomalous subsidence predicted by the dynamical framework. The WTG framework presented here, on the other hand, gives a precise interpretation of how the remote tropical climate responds under El Niño forcing: the climate (and hence anomalous subsidence) responds in such a way that spatial gradients in anomalous tropospheric temperature are negligible. The WTG framework thus naturally lends itself to the nonuniform distribution of anomalous subsidence (as was seen for the observations in Fig. 1), since the level of subsidence consistent with a change to tropospheric temperature in any one vertical column of the QTCM depends on the mean climate of that column, the particular suite of climate feedbacks experienced by the column (with the exception of those processes involving the nondivergent component of the anomalous circulation, which was not modeled), and connections to neighboring columns (through, e.g., moisture fluxes).
Of course, a conceptual framework is only as good as the assumptions underlying it. The simulations outlined in this study demonstrate that while the gross features of the tropical adjustment to ENSO are reasonably simulated—that is, mean tropical tropospheric temperature changes, precipitation responses over the remote Tropics, and the zonal and meridional shifts in convection over the tropical Pacific—the regional-scale spatial features of the remote tropical precipitation and surface temperature anomalies are not well captured. The simulations in which we imposed horizontal tropospheric temperature gradients (section 7) suggest that some of these regional-scale discrepancies arise from the assumption of vanishing tropospheric temperature gradients. Indeed, zonal gradients in tropospheric temperature anomalies are quite apparent in the observational data during ENSO warm phases (cf. Fig. 6 of Yulaeva and Wallace 1994), and the climate impact associated with such gradients may be significant. A determination of the anomalous zonal tropospheric temperature gradients in the standard QTCM simulation compared to a similar simulation in a full 3D atmospheric general circulation model (e.g., CCM3) reveals that the horizontal gradients produced by the QTCM may be somewhat too large, such that the influence of temperature gradients may be overemphasized in the QTCM simulations. In a future study, we aim to investigate the dependence of the tropical ENSO teleconnection on the magnitude of horizontal temperature gradients.
Finally, we briefly comment on the applicability of using El Niño as a proxy for understanding the impacts of global warming, as suggested in previous studies. It is tempting to make an analogy between El Niño– and GHG-induced global warming as both are characterized by increased tropospheric temperatures—indeed, Neelin et al. (2003) and Chou and Neelin (2004) used the warming of the troposphere as a starting point for their upped-ante and rich-get-richer mechanisms for tropical drought regions found in the QTCM under global warming and some El Niño teleconnection regions. Our WTG simulations of the GHG and El Niño scenarios—in which different remote tropical precipitation responses were obtained despite the fact that the tropospheric temperature anomalies were identical in both simulations—suggest that this analogy may be misleading. In particular, we have noted the impact of large-scale, El Niño phase subsidence (induced by anomalous Pacific region convection) on the remote tropical atmospheric humidity field, with subsidence-induced drying of the troposphere during El Niño associated with a reduction of remote-averaged precipitation. In the GHG scenario, where large-scale subsidence does not accompany the temperature increase, the opposite gross-scale precipitation response is observed.
While we do not yet have a complete understanding of the nature of the differences between El Niño and global warming (particularly in terms of causality), it suffices to say that how the tropospheric temperature changes are generated is important—whether temperature changes through the atmospheric greenhouse effect or through energy redistributions by vertical circulations appears to influence the climate response to the forcing. One possible approach to understanding the differences between the El Niño and GHG forcing scenarios is through examination of the moist static energy budget (H. Su 2005, personal communication). In equilibrium, the combination of the moisture and temperature equations leads to a diagnostic equation for vertical pressure velocity as a function of the combined forcing terms for temperature and humidity [the QT and Qq of Eqs. (1) and (2)]. Hence, the difference in the vertical velocity between the GHG and El Niño cases can be tied to the difference of one or more of the individual terms of QT and Qq. Under suitable simplifications—for example, areal averaging and the assumption of weak temperature gradients—it may be possible to associate the differential climate responses to greenhouse gas increases and El Niño to a single-component process (like changes to the top of the atmosphere radiative balance). Such an analysis will be undertaken in a future study.
Acknowledgments
The authors wish to thank Adam Sobel of Columbia University for sharing his expertise and insights into this problem and two anonymous reviewers for providing useful comments and suggestions for improvement of the manuscript. The authors also thank David Neelin and the Climate Systems Interactions Group at UCLA for use of the QTCM and Hui Su and Matt Munnich of UCLA for computational assistance. Funding for this research was provided by NOAA CLIVAR-Pacific Grant NA03OAR4310066.
REFERENCES
Betts, A. K., and M. J. Miller, 1986: A new convective adjustment scheme. 2. Single column tests using gate wave, Bomex, Atex and Arctic air-mass data sets. Quart. J. Roy. Meteor. Soc., 112 , 693–709.
Brown, R. G., and C. S. Bretherton, 1997: A test of the strict quasi-equilibrium theory on long time and space scales. J. Atmos. Sci., 54 , 624–638.
Chiang, J. C. H., and A. H. Sobel, 2002: Tropical tropospheric temperature variations caused by ENSO and their influence on the remote tropical climate. J. Climate, 15 , 2616–2631.
Chiang, J. C. H., and B. R. Lintner, 2005: Mechanisms of remote tropical surface warming during El Niño. J. Climate, 18 , 4130–4149.
Chiang, J. C. H., M. Biasutti, and D. S. Battisti, 2003: Sensitivity of the Atlantic Intertropical Convergence Zone to Last Glacial Maximum boundary conditions. Paleoceanography, 18 .1094, doi:10.1029/2003PA000916.
Chou, C., and J. D. Neelin, 2004: Mechanisms of global warming impacts on regional tropical precipitation. J. Climate, 17 , 2688–2701.
Chou, C., J. D. Neelin, and H. Su, 2001: Ocean–atmosphere–land feedbacks in an idealized monsoon. Quart. J. Roy. Meteor. Soc., 127 , 1869–1891.
Giannini, A., J. C. H. Chang, M. A. Cane, Y. Kushnir, and R. Seager, 2001: The ENSO teleconnection to the tropical Atlantic Ocean: Contributions of the remote and local SSTs to rainfall variability in the tropical Americas. J. Climate, 14 , 4530–4544.
Gill, A. E., 1980: Some simple solutions for heat-induced tropical circulation. Quart. J. Roy. Meteor. Soc., 106 , 447–462.
Hastenrath, S., and L. Heller, 1977: Dynamics of climatic hazards in Northeast Brazil. Quart. J. Roy. Meteor. Soc., 103 , 77–92.
Janicot, S., S. Trzaska, and I. Poccard, 2001: Summer Sahel–ENSO teleconnection and decadal time scale SST variations. Climate Dyn., 18 , 303–320.
Kidson, J. W., 1975: Tropical eigenvector analysis and the Southern Oscillation. Mon. Wea. Rev., 103 , 187–196.
Kiehl, J. T., J. J. Hack, G. B. Bonan, B. A. Boville, D. L. Williamson, and P. J. Rasch, 1998: The National Center for Atmospheric Research Community Climate Model: CCM3. J. Climate, 11 , 1131–1149.
Klein, S. A., B. J. Soden, and N. C. Lau, 1999: Remote sea surface temperature variations during ENSO: Evidence for a tropical atmospheric bridge. J. Climate, 12 , 917–932.
Neelin, J. D., and N. Zeng, 2000: A quasi-equilibrium tropical circulation model—Formulation. J. Atmos. Sci., 57 , 1741–1766.
Neelin, J. D., C. Chou, and H. Su, 2003: Tropical drought regions in global warming and El Niño teleconnections. Geophys. Res. Lett., 30 .2275, doi:10.1029/2003GL018625.
Quinn, W. H., D. O. Zopf, K. S. Short, and R. T. W. K. Yang, 1978: Historical trends and statistics of Southern Oscillation, El Niño, and Indonesian droughts. Fish. Bull., 76 , 663–678.
Ropelewski, C. F., and M. S. Halpert, 1987: Global and regional scale precipitation patterns associated with the El Niño–Southern Oscillation. Mon. Wea. Rev., 115 , 1606–1626.
Saravanan, R., and P. Chang, 2000: Interactions between tropical Atlantic variability and El Niño–Southern Oscillation. J. Climate, 13 , 2177–2194.
Shaevitz, D. A., and A. H. Sobel, 2004: Implementing the weak temperature gradient approximation with full vertical structure. Mon. Wea. Rev., 132 , 662–669.
Sobel, A. H., and C. S. Bretherton, 2000: Modeling tropical precipitation in a single column. J. Climate, 13 , 4378–4392.
Soden, B. J., 1997: Variations in the tropical greenhouse effect during El Niño. J. Climate, 10 , 1050–1055.
Su, H., and J. D. Neelin, 2002: Teleconnection mechanisms for tropical Pacific descent anomalies during El Niño. J. Atmos. Sci., 59 , 2694–2712.
Su, H., J. D. Neelin, and C. Chou, 2001: Tropical teleconnection and local response to SST anomalies during the 1997–1998 El Niño. J. Geophys. Res., 106D , 20025–20043.
Su, H., J. D. Neelin, and J. E. Meyerson, 2004: Tropical tropospheric temperature and precipitation response to sea surface temperature forcing. Ocean–Atmosphere Interaction and Climate Variability, Geophys. Monogr., Vol. 147, Amer. Geophys. Union, 379–392.
Torrence, C., and P. J. Webster, 1999: Interdecadal changes in the ENSO–monsoon system. J. Climate, 12 , 2679–2690.
Webster, P. J., and S. Yang, 1992: Monsoon and ENSO—Selectively interactive systems. Quart. J. Roy. Meteor. Soc., 118 , 877–926.
Wu, Z. X., and R. E. Newell, 1998: Influence of sea surface temperatures on air temperatures in the tropics. Climate Dyn., 14 , 275–290.
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.
Zeng, N., and J. D. Neelin, 1999: A land–atmosphere interaction theory for the tropical deforestation problem. J. Climate, 12 , 857–872.
Zeng, N., J. D. Neelin, and C. Chou, 2000: A quasi-equilibrium tropical circulation model—Implementation and simulation. J. Atmos. Sci., 57 , 1767–1796.

(a) NCEP–NCAR reanalysis 500-mb pressure velocity annual-mean climatology and (b), (c) the differences between El Niño and La Niña phase composite averages and climatology. The annual-mean climatology is computed over 1950–2002. The El Niño (La Niña) phase composite average corresponds to the average annual-mean pressure velocity over all years in which the Niño-3 index is greater (less) than the +1σ (−1σ). In each panel positive (negative) contours are denoted by solid (dashed) lines with negative regions shaded for emphasis. In (a) the contour interval is 1 × 10−2 Pa s−1, while in (b) and (c) the contour interval is 0.25 × 10−2 Pa s−1; contour zero lines are omitted.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

(a) NCEP–NCAR reanalysis 500-mb pressure velocity annual-mean climatology and (b), (c) the differences between El Niño and La Niña phase composite averages and climatology. The annual-mean climatology is computed over 1950–2002. The El Niño (La Niña) phase composite average corresponds to the average annual-mean pressure velocity over all years in which the Niño-3 index is greater (less) than the +1σ (−1σ). In each panel positive (negative) contours are denoted by solid (dashed) lines with negative regions shaded for emphasis. In (a) the contour interval is 1 × 10−2 Pa s−1, while in (b) and (c) the contour interval is 0.25 × 10−2 Pa s−1; contour zero lines are omitted.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1
(a) NCEP–NCAR reanalysis 500-mb pressure velocity annual-mean climatology and (b), (c) the differences between El Niño and La Niña phase composite averages and climatology. The annual-mean climatology is computed over 1950–2002. The El Niño (La Niña) phase composite average corresponds to the average annual-mean pressure velocity over all years in which the Niño-3 index is greater (less) than the +1σ (−1σ). In each panel positive (negative) contours are denoted by solid (dashed) lines with negative regions shaded for emphasis. In (a) the contour interval is 1 × 10−2 Pa s−1, while in (b) and (c) the contour interval is 0.25 × 10−2 Pa s−1; contour zero lines are omitted.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

Precipitation climatologies for the (a) standard and (b) WTG QTCM simulations. The contour interval is 2 mm day−1.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

Precipitation climatologies for the (a) standard and (b) WTG QTCM simulations. The contour interval is 2 mm day−1.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1
Precipitation climatologies for the (a) standard and (b) WTG QTCM simulations. The contour interval is 2 mm day−1.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

El Niño forcing scenario precipitation anomalies. Shown are precipitation anomalies from (a), (b) the standard simulation, (c), (d) the WTG simulation, (e), (f) the standard – WTG precipitation anomaly difference, and (g), (h) the standard – WTG tropospheric temperature anomaly difference. In all panels positive (negative) contours are denoted by solid (dashed) lines with negative contours shaded gray for emphasis. For the Pacific region precipitation and precipitation difference anomalies, in (a), (c), and (e), a contour interval of 1 mm day−1 is used; while for the remote region precipitation and precipitation difference anomalies, in (b), (d), and (f), a contour interval of 0.25 mm day−1 is used. In (g) and (h) a contour interval of 0.1 K is used.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

El Niño forcing scenario precipitation anomalies. Shown are precipitation anomalies from (a), (b) the standard simulation, (c), (d) the WTG simulation, (e), (f) the standard – WTG precipitation anomaly difference, and (g), (h) the standard – WTG tropospheric temperature anomaly difference. In all panels positive (negative) contours are denoted by solid (dashed) lines with negative contours shaded gray for emphasis. For the Pacific region precipitation and precipitation difference anomalies, in (a), (c), and (e), a contour interval of 1 mm day−1 is used; while for the remote region precipitation and precipitation difference anomalies, in (b), (d), and (f), a contour interval of 0.25 mm day−1 is used. In (g) and (h) a contour interval of 0.1 K is used.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1
El Niño forcing scenario precipitation anomalies. Shown are precipitation anomalies from (a), (b) the standard simulation, (c), (d) the WTG simulation, (e), (f) the standard – WTG precipitation anomaly difference, and (g), (h) the standard – WTG tropospheric temperature anomaly difference. In all panels positive (negative) contours are denoted by solid (dashed) lines with negative contours shaded gray for emphasis. For the Pacific region precipitation and precipitation difference anomalies, in (a), (c), and (e), a contour interval of 1 mm day−1 is used; while for the remote region precipitation and precipitation difference anomalies, in (b), (d), and (f), a contour interval of 0.25 mm day−1 is used. In (g) and (h) a contour interval of 0.1 K is used.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

As in Fig. 3 but for La Niña forcing scenario.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

As in Fig. 3 but for La Niña forcing scenario.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1
As in Fig. 3 but for La Niña forcing scenario.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

Surface temperature anomalies for the (a) standard and (b) WTG QTCM and (c) the standard – WTG surface temperature difference. In (a) and (b) the grayscale key at the right denotes anomalies in units of K. In (c) positive (negative) contours are denoted by solid (dashed) lines, and the contour interval is 0.1 K, with negative contours shaded gray for emphasis.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

Surface temperature anomalies for the (a) standard and (b) WTG QTCM and (c) the standard – WTG surface temperature difference. In (a) and (b) the grayscale key at the right denotes anomalies in units of K. In (c) positive (negative) contours are denoted by solid (dashed) lines, and the contour interval is 0.1 K, with negative contours shaded gray for emphasis.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1
Surface temperature anomalies for the (a) standard and (b) WTG QTCM and (c) the standard – WTG surface temperature difference. In (a) and (b) the grayscale key at the right denotes anomalies in units of K. In (c) positive (negative) contours are denoted by solid (dashed) lines, and the contour interval is 0.1 K, with negative contours shaded gray for emphasis.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

As in Fig. 3 but for ΔGHG+ forcing scenario.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

As in Fig. 3 but for ΔGHG+ forcing scenario.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1
As in Fig. 3 but for ΔGHG+ forcing scenario.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

Histograms of precipitation anomalies vs mean precipitation values in the (a) standard and (b) WTG QTCM. Solid (dashed) lines represent results of the El Niño (ΔGHG+) forcing scenarios. Each curve represents the average over the remote tropical region (20°S–20°N, 70°W–110°E) for the last 36 months of each simulation. The error bars represent the 2σ level of the bin average, with bin widths taken as 2 mm day−1.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

Histograms of precipitation anomalies vs mean precipitation values in the (a) standard and (b) WTG QTCM. Solid (dashed) lines represent results of the El Niño (ΔGHG+) forcing scenarios. Each curve represents the average over the remote tropical region (20°S–20°N, 70°W–110°E) for the last 36 months of each simulation. The error bars represent the 2σ level of the bin average, with bin widths taken as 2 mm day−1.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1
Histograms of precipitation anomalies vs mean precipitation values in the (a) standard and (b) WTG QTCM. Solid (dashed) lines represent results of the El Niño (ΔGHG+) forcing scenarios. Each curve represents the average over the remote tropical region (20°S–20°N, 70°W–110°E) for the last 36 months of each simulation. The error bars represent the 2σ level of the bin average, with bin widths taken as 2 mm day−1.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

Vertically averaged tropical humidity anomalies as a function of longitude. The results shown are 20°S–20°N averages of vertically averaged tropospheric humidity (in units of K) for the WTG El Niño (squares) and WTG ΔGHG+ (triangles) simulations.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

Vertically averaged tropical humidity anomalies as a function of longitude. The results shown are 20°S–20°N averages of vertically averaged tropospheric humidity (in units of K) for the WTG El Niño (squares) and WTG ΔGHG+ (triangles) simulations.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1
Vertically averaged tropical humidity anomalies as a function of longitude. The results shown are 20°S–20°N averages of vertically averaged tropospheric humidity (in units of K) for the WTG El Niño (squares) and WTG ΔGHG+ (triangles) simulations.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

Precipitation response for the WTG simulation with anomalous temperature gradients. Positive (negative) contours are denoted by solid (dashed) lines with negative contours shaded gray for emphasis. (a) For the Pacific region precipitation anomalies, a contour interval of 1 mm day−1 is used, and (b) for the remote region precipitation anomalies, a contour interval of 0.25 mm day−1 is used.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1

Precipitation response for the WTG simulation with anomalous temperature gradients. Positive (negative) contours are denoted by solid (dashed) lines with negative contours shaded gray for emphasis. (a) For the Pacific region precipitation anomalies, a contour interval of 1 mm day−1 is used, and (b) for the remote region precipitation anomalies, a contour interval of 0.25 mm day−1 is used.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1
Precipitation response for the WTG simulation with anomalous temperature gradients. Positive (negative) contours are denoted by solid (dashed) lines with negative contours shaded gray for emphasis. (a) For the Pacific region precipitation anomalies, a contour interval of 1 mm day−1 is used, and (b) for the remote region precipitation anomalies, a contour interval of 0.25 mm day−1 is used.
Citation: Journal of Climate 18, 24; 10.1175/JCLI3580.1
Comparison of area-averaged tropical 500-mb annual-mean pressure velocity values. Averages are computed from the 500-mb NCEP–NCAR reanalysis vertical velocity field for meridional domains of 20°S–20°N and 30°S–30°N spanning the Pacific (110°E–70°W), remote Tropics (70°W–110°E), and entire Tropics. The results represent time averages over all years between 1950 and 2002 as well as those years corresponding to either El Niño or La Niña phase conditions. El Niño (La Niña) phase conditions are defined as those years in which the annual-mean Niño-3 index is greater than (less than) + (−) one standard deviation, with the standard deviation computed over 1950–2002. Values are in units of Pa s−1 (× 10−2), and negative (positive) values correspond to upward (downward) motion.


Select climate parameters averaged over the remote region for the standard and WTG QTCM El Niño and La Niña phase simulations. Tabulated values are averaged over 20°S–20°N, 70°W–110°E for the last 3 years of the simulation. Note that specific humidity is given in units of K; to convert to more conventional units (i.e., g kg−1), divide the value in K by 2.42.

