The Tropical Atmospheric El Niño Signal in Satellite Precipitation Data and a Global Climate Model

Yonghua Chen Department of Applied Physics and Applied Mathematics, Columbia University, and NASA Goddard Institute for Space Studies, New York, New York

Search for other papers by Yonghua Chen in
Current site
Google Scholar
PubMed
Close
,
Anthony D. Del Genio NASA Goddard Institute for Space Studies, New York, New York

Search for other papers by Anthony D. Del Genio in
Current site
Google Scholar
PubMed
Close
, and
Junye Chen Earth System Science Interdisciplinary Center, University of Maryland, College Park, College Park, and Global Modeling and Assimilation Office, NASA Goddard Space Flight Center, Greenbelt, Maryland

Search for other papers by Junye Chen in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

Aspects of the tropical atmospheric response to El Niño related to the global energy and water cycle are examined using satellite retrievals from the Tropical Rainfall Measuring Mission and the Advanced Microwave Scanning Radiometer-E and simulations from the Goddard Institute for Space Studies (GISS) general circulation model (GCM). The El Niño signal is extracted from climate fields using a linear cross-correlation technique that captures local and remote in-phase and lagged responses. Passive microwave and radar precipitation anomalies for the 1997/98 and 2002/03 El Niños and the intervening La Niña are highly correlated, but anomalies in stratiform–convective rainfall partitioning in the two datasets are not. The GISS GCM produces too much rainfall in general over ocean and too little over land. Its atmospheric response to El Niño is weaker and decays a season too early. Underestimated stratiform rainfall fraction (SRF) and convective downdraft mass flux in the GISS GCM and excessive shallow convective and low stratiform cloud result in latent heating that peaks at lower altitudes than inferred from the data. The GISS GCM also underestimates the column water vapor content throughout the Tropics, which causes it to overestimate outgoing longwave radiation. The response of both quantities to interannual Hadley circulation anomalies is too weak. The GISS GCM’s Walker circulation also exhibits a weak remote response to El Niño, especially over the Maritime Continent and western Indian Ocean. This appears to be a consequence of weak static stability due to the model’s lack of upper-level stratiform anvil heating, excessive low-level heating, and excessive dissipation due to cumulus momentum mixing. Our results suggest that parameterizations of mesoscale updrafts, convective downdrafts, and cumulus-scale pressure gradient effects on momentum transport are keys to a reasonable GISS GCM simulation of tropical interannual variability.

Corresponding author address: Anthony D. Del Genio, NASA Goddard Institute for Space Studies, 2880 Broadway, New York, NY 10025. Email: adelgenio@giss.nasa.gov

Abstract

Aspects of the tropical atmospheric response to El Niño related to the global energy and water cycle are examined using satellite retrievals from the Tropical Rainfall Measuring Mission and the Advanced Microwave Scanning Radiometer-E and simulations from the Goddard Institute for Space Studies (GISS) general circulation model (GCM). The El Niño signal is extracted from climate fields using a linear cross-correlation technique that captures local and remote in-phase and lagged responses. Passive microwave and radar precipitation anomalies for the 1997/98 and 2002/03 El Niños and the intervening La Niña are highly correlated, but anomalies in stratiform–convective rainfall partitioning in the two datasets are not. The GISS GCM produces too much rainfall in general over ocean and too little over land. Its atmospheric response to El Niño is weaker and decays a season too early. Underestimated stratiform rainfall fraction (SRF) and convective downdraft mass flux in the GISS GCM and excessive shallow convective and low stratiform cloud result in latent heating that peaks at lower altitudes than inferred from the data. The GISS GCM also underestimates the column water vapor content throughout the Tropics, which causes it to overestimate outgoing longwave radiation. The response of both quantities to interannual Hadley circulation anomalies is too weak. The GISS GCM’s Walker circulation also exhibits a weak remote response to El Niño, especially over the Maritime Continent and western Indian Ocean. This appears to be a consequence of weak static stability due to the model’s lack of upper-level stratiform anvil heating, excessive low-level heating, and excessive dissipation due to cumulus momentum mixing. Our results suggest that parameterizations of mesoscale updrafts, convective downdrafts, and cumulus-scale pressure gradient effects on momentum transport are keys to a reasonable GISS GCM simulation of tropical interannual variability.

Corresponding author address: Anthony D. Del Genio, NASA Goddard Institute for Space Studies, 2880 Broadway, New York, NY 10025. Email: adelgenio@giss.nasa.gov

1. Introduction

El Niño–Southern Oscillation (ENSO) is the most prominent mode of unforced interannual variability in the global climate system. As such it is considered to be a good test for evaluating climate model performance. It is not directly diagnostic of long-term climate feedbacks, since El Niño sea surface temperature (SST) anomaly patterns, and the tropospheric circulation response they induce, are likely to differ from those in a forced climate change (Del Genio 2002). However, the ability to simulate the atmospheric ENSO signal in hydrological cycle fields is a necessary if not sufficient condition for having confidence in a model’s regional and global predictions of climate change.

Current general circulation models (GCMs) underestimate atmospheric responses to observed ENSO SST anomalies, especially in fields associated with the global energy and water cycle (e.g., Soden 2000). Precipitation, as one of the most important climate variables, has been a challenge for climate models to simulate. Dai (2006) indicates that the simulations of precipitation in the newest generation of coupled atmosphere–ocean GCMs still need considerable improvement in their temporal and spatial variability, precipitation frequency, and stratiform–convective partitioning. Examining a GCM’s precipitation field in detail in the context of the whole hydrological cycle response to ENSO perturbations will hopefully provide clues to areas of the model’s cumulus parameterization that need further development and improvement.

State-of-the-art satellite observations during the last two decades now provide a record sufficiently long to document several El Niño and La Niña episodes and the corresponding atmospheric responses. The Tropical Rainfall Measuring Mission (TRMM) satellite and more recent Earth Observing System (EOS) satellites observe a variety of parameters related to the atmospheric hydrological response to ENSO, such as precipitation, latent heating, and column water vapor, during the 1997/98 and 2002/03 El Niño events. TRMM covers the entire Tropics with good diurnal sampling, while the EOS satellites each observe twice daily but cover latitudes up to ±70°.

The goals of this paper are threefold. First, we supplement and extend the existing observational record of ENSO variations in precipitation (Yulaeva and Wallace 1994; Dai et al. 1997; Huffman et al. 1997; Xie and Arkin 1998; Curtis and Adler 2000; Robertson et al. 2001; Berg et al. 2002), radiation budget (Cess et al. 2001; Lu et al. 2004), and water vapor (Bates et al. 2001). Second, given the sampling and retrieval uncertainties in satellite rainfall observations, we examine the consistency in the ENSO signals of multiple precipitation datasets. Third, we explore the realism of the Goddard Institute for Space Studies (GISS) GCM’s atmospheric response to El Niño, and possible parameterization reasons for its deficiencies in this area. The paper is organized as follows: in the next three sections, we describe the observations, the GCM used, and the extraction of the ENSO signal; in section 5 simulated hydrological and dynamical variables are compared with the observations in detail; and the last section summarizes the evaluation and discusses possible reasons for the discrepancies found in this study.

2. Data sources

a. Rainfall products

Three monthly gridded products (version 6) from the TRMM satellite (Kummerow et al. 2000), covering the period December 1997–June 2005, are included in this study: TRMM 3A12 (Kummerow et al. 2001) from the TRMM Microwave Imager (TMI), TRMM 3A25–6A (Iguchi et al. 2000) from the Precipitation Radar (PR), and TRMM convective–stratiform heating (CSH; Tao et al. 2001) from the PR. TMI data include surface total rain rate, stratiform–convective partitioning, and latent heating profiles at 0.5° × 0.5° resolution covering ±40° latitude. PR data include surface total rain rate and stratiform–convective partitioning at 0.5° × 0.5° resolution with similar area coverage as TMI but fewer samples per unit area due to its narrower swath. Although these two products are based on two different data sources and retrieval algorithms, their rainfall rates are highly correlated, with TMI values being slightly higher than those from PR (Kummerow et al. 2000). CSH provides monthly 0.5° × 0.5° latent heating profiles derived from surface convective and stratiform rain rates.

Several caveats about the TRMM rainfall rates should be noted. Tropical mean ENSO rainfall anomalies differed significantly in earlier versions of the TMI and PR data (Berg et al. 2002; Robertson et al. 2003), but discrepancies have been reduced in version 6. In addition, the TRMM satellite’s orbit was boosted in August 2001 to extend its lifetime. The larger footprints decrease the detection of weak rainfall and thus increase the PR conditional mean rain rate slightly (Shimizu et al. 2003), while the effect on TMI appears to be a more substantial increase in high rain-rate areas (De Moss and Bowman 2007). Since we focus only on regional components of change that correlate with an ENSO SST index, we expect these problems to have little effect on our results.

The retrieval methods of rain type (convective versus stratiform) for TMI and PR are different as well. The TMI retrievals are indirect, based on texture in the horizontal brightness temperature gradient and the 85.5-GHz polarization signatures (Olson et al. 2001). The PR retrievals are based on the radar reflectivity profile, in which a bright band indicates the stratiform region (Awaka et al. 1997). This is a direct physically based method, though the horizontal texture of the reflectivity field is also used when ambiguity exists. The separation of convective and stratiform precipitation plays an important role in correctly estimating the latent heating profiles in the Tropics (Houze 1982, 1997): convective heating is concentrated in the lower troposphere, where water vapor concentrations and thus condensation rates in cumulus updrafts are greatest, while stratiform heating peaks in the middle and upper troposphere where anvils form above the freezing level. The TMI-based latent heating profiles are derived using a technique described by Olson et al. (1999, 2006). The CSH algorithm of Tao et al. (1993) utilizes the stratiform fraction of rain and lookup tables simulated from a cloud-resolving model or estimated from a diagnostic budget study.

The Advanced Microwave Scanning Radiometer–Earth Observing System (AMSR-E) instrument on the EOS Aqua satellite provides passive microwave measurements of terrestrial, oceanic, and atmospheric parameters twice daily covering up to ±70° latitude. The level-3 rainfall accumulation product (AE_RnGd) used in this study is the monthly averaged rainfall accumulation over ocean and land at 5° × 5° resolution for the period June 2002–June 2005 (Wilheit et al. 2003; Adler et al. 2004).

b. Other TRMM products

Other atmospheric parameters over ocean from TMI are available from Remote Sensing Systems (RSS) online (see http://www.ssmi.com/tmi/tmi_browse.html). The monthly gridded TMI ocean product (version 3a) includes SST, column water vapor (CWV), and surface wind speed (Wentz and Meissner 2000). The resolution is 0.25° × 0.25° for the same area and temporal coverage as TRMM 3A12.

c. Other datasets

Monthly mean vertical profiles of pressure vertical velocity (omega) and temperature through June 2005 are obtained from the National Centers for Environmental Prediction–National Center for Atmospheric Research (NCEP–NCAR) reanalysis (Kalnay et al. 1996) at 2.5° × 2.5° resolution. Outgoing longwave radiation (OLR) at the top of the atmosphere (TOA) through December 2004 is obtained from the International Satellite Cloud Climatology Project radiative flux dataset (ISCCP-FD; Zhang et al. 2004).

3. GISS global climate model description

We use an updated SI2000 version of the GISS general circulation model (GCM), but at higher horizontal and vertical resolutions (2° × 2.5° × 32L) than the previous version documented by Hansen et al. (2002). Most model physics has been updated to match that in the more recent Model E GCM used for long-term climate change experiments (Schmidt et al. 2006). Relevant features of the model for this study include the following. Advection of water vapor is based on a quadratic upstream scheme. The radiation uses the correlated-k distribution method and a single Gauss point adaptation of the doubling–adding method. The boundary layer is based on a nonlocal turbulence scheme using dry rather than moist conserved variables.

The GCM produces convective precipitation using the bulk mass flux scheme of Del Genio and Yao (1993), and stratiform precipitation from the prognostic cloud water parameterization of Del Genio et al. (1996). The cumulus parameterization scheme uses a closure assumption based on a stability-dependent mass flux. Triggering is based on the local buoyancy of parcels lifted between two model layers, with a mass flux that results in neutral buoyancy at cloud base. The mass flux spectrum is simulated by two plumes: one undilute and the other entraining at a fixed rate. Detrainment occurs at the level of neutral buoyancy. Convective-scale downdrafts are created from negatively buoyant equal mixtures of updraft and environmental air, the mass flux specified to be a third that of the updrafts. Convective motions carry momentum from their level of origin.

For large-scale cloud formation, the nonconvective portion of a grid box is divided into cloudy and clear parts based on relative humidity. A single cloud water variable is predicted with a condensation source based on available moisture convergence. Simple representations of all microphysical processes (autoconversion, accretion, Bergeron–Findeisen diffusional growth, evaporation, and cloud-top entrainment) are included. Cloud phase is based on a temperature-dependent probability and adjusted when glaciation of supercooled water by falling ice is diagnosed. Cloud optical thickness is diagnosed from the cloud water path assuming different fixed particle number concentrations for continental liquid, ocean liquid, and ice clouds. Of particular relevance for this study is the coupling between the convection and stratiform cloud schemes. Cumulus updraft condensate follows a Marshall–Palmer distribution. Empirical size–fall speed relationships and assumed updraft speed profiles are used to partition the drop-size distribution into a precipitating fraction and a remainder that detrains to form an anvil (Del Genio et al. 2005). Subsequent anvil evolution depends on the parameterized stratiform cloud microphysics, but mesoscale updrafts and downdrafts are not included.

The model simulation is done with atmospheric components only but forced by observed SST and sea ice cover from the Hadley Centre Sea Ice and SST dataset (HadISST1; Rayner et al. 2003) for the period November 1996 to June 2005. The various datasets and the GCM have different resolutions, so interpolation is applied in order to match the lowest resolution for all analyses.

4. Methodology

To study the atmospheric response to El Niño, the first step is to characterize the ENSO event in the observations and the GCM. Numerous methods have been used previously to define the ENSO anomaly in the surface temperature field (e.g., Kelly and Jones 1996; Santer et al. 2001). However, ENSO affects surface temperature as well as other climate variables with different time lags in different regions globally (Lanzante 1996). Using a single lag–lead relationship to obtain ENSO-related variability for all regions is problematic (e.g., Robock and Mao 1995). In this study, we use a simple but effective lag–correlation technique to extract the ENSO signal from a given monthly mean field at the grid box level. This method, developed by Chen (2005), objectively defines the ENSO anomaly and captures both the in-phase and the lagged local and remote response. Details of the procedure are described in the appendix.

We apply the technique to the TMI SST data and demonstrate its ability to capture known features of the ENSO evolution of SST (Klein et al. 1999) and corresponding lag–lead relationships revealed by other techniques such as complex principal component analysis (Lanzante 1996). ENSO events can be classified by their strength and SST anomaly pattern. Fu et al. (1986) define type-I events as those with significant warming in the central and east Pacific, and type-II events as those whose warming extends farther westward and is weaker in the eastern Pacific. The different patterns of warming can result in different regional remote responses. The peak and decline of the strong type-I 1997/98 El Niño are clearly shown in Fig. 1: a positive SST anomaly expanding from the date line to the east Pacific, and a negative SST anomaly over the west Pacific and the subtropics in both hemispheres; the anomalies begin to decay in boreal spring 1998. A La Niña appears in boreal autumn 1998 with a negative SST anomaly over the central-eastern Pacific and a positive anomaly over the west Pacific, persisting through boreal winter 2001. The technique also captures the moderate 2002/03 El Niño, which starts in the west Pacific and is more type II in structure. This event decays in mid-2003 but reappears as a third weak event in late 2004. The remote lagged response in the Indian Ocean (IO; cf. Lanzante 1996) can also be seen for each event with a positive SST anomaly during El Niño and a negative anomaly during La Niña.

Other aspects of the in-phase and the lag–lead relationships on different time scales in different regions can be seen in Fig. 2. SST anomalies throughout the central/east equatorial Pacific, extending to the eastern subtropics, are highly positively correlated with the Niño-3.4 index (N34). The correlation is in phase near the equator but lags N34 by several months in the subtropics. The equatorial west Pacific and remaining regions of the subtropical Pacific are negatively correlated with N34 at lags of several months, consistent with the time required for circulation anomalies to remotely influence the surface energy budget. In the Indian Ocean, the SST anomaly lags N34 by about 1∼5 months; this may also be due to remote ENSO forcing of circulation and surface heat flux changes (Klein et al. 1999), but it may also be associated with a separate Indian Ocean dipole (Saji et al. 1999), which may in turn influence ENSO evolution (Kug and Kang 2006). Several Pacific locations appear to lead N34, but these tend to occur near the transition from positive to negative correlation where the ENSO signal is weak.

5. Results

In this section, we first compare the various precipitation retrievals to each other and to the model to characterize the observational uncertainties and define the strengths and weaknesses of the GCM’s precipitation climatology and ENSO signal. We then analyze higher-order diagnostics of the precipitation field such as stratiform–convective partitioning, latent heating profiles, and related hydrological and dynamical fields, to gain insight into possible causes of the GCM deficiencies.

a. Precipitation

The annual mean precipitation distribution from TMI, PR, and the GCM is plotted in Fig. 3. There are some discrepancies between the TMI and PR annual mean precipitation both in the spatial patterns and zonal mean distributions: TMI rain rates generally exceed those of PR, and TMI maxima over central Africa, the Himalayas, and the Amazon basin are missing or only weakly present in PR, which is more reliable over land. But in general, the differences are small compared to those between the GCM and either dataset. The simulated Pacific intertropical convergence zone (ITCZ) and South Pacific convergence zone (SPCZ) are too broad and their rainfall rates too intense. There is also an unrealistically strong rainfall maximum in the tropical west Atlantic and a spurious maximum over the north Indian Ocean. The precipitation maximum over the Amazon is underestimated, while a spurious peak over the Himalayas not seen by PR is simulated. The seasonal mean precipitation (not shown) has similar patterns of deficiency in the GCM in that there is stronger rainfall than observed over most ocean areas and weaker over some land areas. Over the Maritime Continent (MC), the GCM produces more (less) rainfall in the north (south) part during boreal winter and spring, and behaves in the opposite sense during boreal summer and fall.

Figure 4 compares the evolution of seasonal ENSO anomalies in precipitation for TMI (Fig. 4a) and the GCM (Fig. 4b). The ENSO-induced precipitation patterns are characterized by a positive anomaly over the central and eastern Pacific, and a negative anomaly over the western Pacific and the MC during El Niño, and vice versa during La Niña (Fig. 4a). Positive Pacific rainfall anomalies are less extensive than SST anomalies, however. Rainfall decreases during the warm phase in the climatological ITCZ and SPCZ locations, including the east Pacific ITCZ where SST warms. Positive IO rainfall anomalies occur during the warm phase, but only in the western part of the basin, while SST warms over most of the IO. Although the GCM simulation shows a similar pattern of El Niño– and La Niña–like anomalies, the magnitude is weaker than observed in most areas (Fig. 4b). Furthermore, the anomaly pattern is more zonally oriented than observed, suggesting that the model’s Walker cell perturbation is underestimated, especially over the MC and IO regions. The GCM also misses the observed precipitation changes in the southeastern United States during the 1997/98 ENSO event. Finally, tropical anomalies of both signs generally weaken a season too early compared to those observed, for example, MAM in 1998, 2000, and 2003 in Fig. 4b.

Figures 5a,b show the correlation between TMI and PR monthly gridbox precipitation anomalies within the dataset domain for El Niño (N34 > 0.5) and La Niña (N34 < −0.5) periods over the period December 1997–June 2005. The two datasets are highly correlated (correlation coefficient R = 0.9) but with PR anomalies only 80%–85% as strong as TMI anomalies. The largest differences typically occur in the central equatorial Pacific. For the June 2002–June 2005 period that overlaps AMSR-E, which contained only a single weak ENSO event, we use positive and negative values of N34 instead to separate El Niño and La Niña. The resulting PR–TMI correlation (Figs. 5c,d) is lower (R = 0.7), and the PR signal only about 2/3 the strength of the TMI signal. In part this may be due to the ambiguity in isolating ENSO over the shorter time interval, but it may also reflect the greater disparity in the two products at weaker rain rates where TMI and PR disagree the most (Kummerow et al. 2001), since this period contained only moderate ENSO anomalies. The GCM (not shown), whose ENSO anomalies are displaced relative to those observed in some locations, is correlated more weakly with TMI (R = 0.5), and its anomalies are only ∼40% as large.

Figures 5e,f compare AMSR-E and TMI anomalies over this same period. The strengths of AMSR-E and TMI ENSO anomalies are more similar than those of PR and TMI, though the correlation is slightly worse. The latter is due primarily to a small number of grid boxes, mostly over the Himalayas, where the AMSR-E monthly product using the Wilheit et al. (2003) algorithm diagnoses large anomalies while the TMI algorithm of Kummerow et al. (2001) does not. Excluding these, Figs. 5e,f suggest that the sparse diurnal sampling of AMSR-E does not seriously degrade its ENSO precipitation signal.

Figure 6 shows the extreme correlation at least lag (see appendix) of precipitation with N34 and corresponding lag–lead relationship maps for PR (Fig. 6a), TMI (Fig. 6b), and the GCM (Fig. 6c). Both TMI and PR are negatively correlated with N34 over the eastern IO, the MC, the SPCZ, the Amazon basin, and the west equatorial Atlantic Ocean, and positively correlated over the central–east equatorial Pacific Ocean (upper panels in Fig. 6a and Fig. 6b). The local central–east Pacific precipitation anomaly tends to lag the evolution of Pacific SST by 1–2 months except right at the equator (lower panels in Fig. 6a and Fig. 6b), suggesting that low-level moisture convergence anomalies must develop before precipitation is affected. The observed precipitation anomaly in the MC region leads the Pacific SST anomaly, consistent with the finding of Curtis and Adler (2000). This implies that subsidence anomalies due to weakening trade winds may determine MC behavior as much as the delayed remote response to resulting east Pacific SST anomalies. The GCM simulates reasonably well the observed positive correlation between equatorial Pacific precipitation anomalies and SST, and the negative correlation in the SPCZ and Atlantic. However, the strong negative (positive) correlations observed in the MC (IO) regions are missing (misplaced) in the model. The missing MC response, if driven by trade wind rather than SST anomalies at ENSO initiation, may be a limitation of atmosphere-only GCM simulations in which the surface wind can respond to but not affect the prescribed SST.

b. Stratiform–convective partitioning and latent heating

The release of latent heat in precipitating areas is a major driving force to the large-scale circulation in the Tropics. The profile of latent heating is associated with the partitioning of stratiform versus convective rainfall (Houze 1982, 1997). In a convective dominated region, heating peaks in the lower troposphere, while it is concentrated in the upper troposphere (with evaporative cooling below) in the stratiform anvil region. Changes in the stratiform rainfall fraction (SRF) and corresponding latent heating profile during El Niño events have been reported by Schumacher and Houze (2003), who show an enhanced near-equatorial trans-Pacific SRF gradient during El Niño.

The two TRMM SRF products show different features both in the mean and ENSO anomaly fields. For the mean field (Fig. 7, left), the largest discrepancies are found over land and over the east Pacific ITCZ. (Large apparent PR–TMI discrepancies over the Peruvian and Namibian marine stratocumulus regions are not associated with deep convective systems and are irrelevant to this discussion.) In general, PR stratiform fraction over ocean is ∼40%–60% and that for TMI is ∼10% smaller in heavily precipitating regions. Over land, the PR SRF is somewhat smaller and the TMI SRF is considerably smaller. The GCM simulation has a general pattern that agrees with PR, but with much larger SRF than PR and TMI have over land (where GCM mean rainfall is deficient), and much smaller SRF over most ocean areas (∼10%–20%), symptomatic of the absence of a mesoscale updraft parameterization in the model (Donner et al. 2001). Corresponding to those differences in SRF between the GCM and data over ocean regions, the GCM produces more convective precipitation, and thus latent heating peaks at a much lower altitude with a much stronger magnitude compared to the TRMM latent heating profile (see below and section 6).

In the right column of Fig. 7, the 1998 December–February (DJF) SRF ENSO anomaly is plotted for PR (Fig. 7d), TMI (Fig. 7e), and the GCM (Fig. 7f). The sign of the SRF ENSO anomaly is opposite in TMI and PR data. PR has a negative SRF anomaly over the MC region, the SPCZ, and the subtropical Pacific in the Northern Hemisphere, and a strong positive anomaly over the central/east Pacific (Schumacher and Houze 2003). TMI instead shows a positive 1997/98 ENSO anomaly in the subtropical Pacific in both hemispheres, and weak negative changes over the MC region and the central equatorial Pacific Ocean. Similar patterns exist during the 2002/03 El Niño (not shown), but the area covered and anomaly magnitude are smaller. The GCM’s SRF ENSO anomaly pattern resembles the less reliable TMI estimate but is stronger over the central Pacific.

Mean and ENSO 1998 DJF anomalies in the altitude of peak latent heating (ALPH) for the CSH product and the GCM are shown in Fig. 8. The GCM plots are shown for the deep convective component of heating only, since shallow nonprecipitating clouds are not detected by PR and do not contribute to the CSH profiles. The PR retrieval peaks in the middle troposphere in convecting regions, higher in the west Pacific warm pool and over the tropical continents than in the east Pacific ITCZ. ENSO anomalies in the PR-based ALPH are fairly consistent with the PR SRF anomalies in the central Pacific (ALPH rises where SRF increases during the warm phase), the SPCZ, and the MC (ALPH and SRF both decrease). We also examined the TMI-based latent heating profiles (not shown), which show a higher ALPH over most ocean areas than the PR-based product. The TMI ENSO perturbation latent heating is also noisier than the PR anomaly. The patterns in the GCM’s ALPH are similar to those in CSH, but its latent heating peaks at a lower altitude than that of CSH over the oceans and most continental regions, and at higher altitudes over a few tropical land locations. The GCM’s ALPH also increases in the central/east Pacific during the warm phase despite its SRF decreasing there, suggesting that it is determined by an upward shift in the convective heating profile as rain rate increases rather than a change in the weighting between stratiform and convective contributions. The big decrease of ALPH during El Niño in the MC, however, is missing from the GCM.

The DJF latent heating profiles from shallow convective, deep convective, and stratiform clouds in the GCM are shown in Fig. 9. The deep convective heating peaks near 600 mb but has significant heating down to 800 mb. The latter feature is consistent with the GCM’s underestimated downdraft mass flux (Xie et al. 2002). Shallow convection and heating from low-level stratiform cloud further depress the level of the overall heating. The upper-level stratiform heating shows the effect of the absent mesoscale updraft—rather than a strong upper-troposphere heating peak with evaporative cooling below, the model exhibits only slight net heating above the 300-mb level and a deep layer of weak evaporative cooling below that.

c. Other hydrological fields

In this subsection, we explore the atmospheric responses to El Niño in several other components of the tropical water and energy cycle: CWV, OLR (upward positive), and middle-tropospheric temperature (TT). In the subtropical descending branch of the Hadley cell, CWV is the primary influence on OLR and is also an indicator of the strength of the descending motion. In the ITCZ, OLR is controlled by cumulus anvil and thinner cirrus clouds and is thus diagnostic of the ice water path produced by convection and large-scale ascent. The warming of tropospheric temperature and its spread across the entire tropical belt via propagating Kelvin and Rossby waves has been proposed as an ENSO teleconnection mechanism for remote tropical surface warming (Chiang and Lintner 2005).

The GCM CWV mean field is almost uniformly lower than that observed by TMI (by ∼12%) throughout the Tropics (not shown). This corresponds to a GCM OLR mean value larger than the ISCCP mean (by ∼9.5 W m−2). The mean middle-tropospheric temperature at 400 mb is reasonably simulated in the GCM. Over the Pacific and Indian Oceans, the CWV anomaly has the same sign and similar patterns as precipitation, while the OLR anomaly has the opposite sign but similar patterns (Figs. 10a,b). We are especially interested in the CWV and OLR ENSO signal over the subtropical subsidence regions (i.e., 15°–30°S and 15°–30°N). Figure 11 shows scatterplots of the monthly 500-mb omega anomaly versus CWV anomaly for December 1997 to June 2005 from the NCEP–NCAR reanalysis and TMI (top panel) and the GCM (bottom panel). The response of the model’s subtropical CWV to ENSO Hadley circulation anomalies is too weak by ∼40%. This may be due to its underestimate of SRF, which weakens the remote circulation response (Schumacher et al. 2004) and thus subsidence drying. The model’s weaker-than-observed negative OLR anomaly in the central–east equatorial Pacific in Fig. 10b is diagnostic of an underestimated ENSO increase in cumulus condensate detrainment and/or anvil ice production, also consistent with the model’s SRF problem. Details of the vertical structure of the response over different regions will be discussed in section 6.

The ENSO-induced anomaly in NCEP–NCAR reanalysis and GCM TT are compared in Fig. 10c. The signal is dominated by a TT warming response to El Niño through most of the tropical belt both in NCEP–NCAR and the GCM, consistent with previous studies (e.g., Yulaeva and Wallace 1994). NCEP–NCAR shows the classic Gill (1980)-type pattern of response to a surface heating anomaly. The initial eastern equatorial Pacific warming generates a Kelvin wave that propagates to the east, while Rossby waves on either side of the equator propagate the signal to the west Pacific. This is proposed by Chiang and Sobel (2002) and Chiang and Lintner (2005) as a mechanism for ENSO impacts in the remote Tropics. Although the GISS GCM produces TT warming across the Pacific, the Gill pattern is less obvious. The Kelvin wave remote influence over Africa and the west IO is weaker. Some features of the Rossby wave pattern do not propagate sufficiently far westward—the primary warm anomaly barely extends past the date line, and cooling anomalies over east Asia and west Australia–southeast IO are mostly missing in the GCM.

6. Summary and discussion

The results of the previous section can be summarized as follows. Although the GCM produces a generally reasonable precipitation ENSO anomaly pattern with decreased rainfall in the west Pacific and increased rainfall in the central–eastern Pacific during El Niño, which agrees with TRMM data, it has difficulty in simulating the remote El Niño precipitation responses over the Maritime Continent, the Indian Ocean, and the southeastern United States. Stratiform rainfall is underestimated in the GCM over the tropical oceans both in the mean and ENSO anomaly fields. This specific feature of the GISS model may be generally relevant to climate GCMs: Dai (2006) shows that most of the 18 state-of-the-art coupled atmosphere–ocean GCMs produce too much convective and too little stratiform rainfall over the Tropics. This causes a weak upper-tropospheric latent heating response. Thus, moist static energy export from convecting regions is likely to be too weak, possibly accounting for the GCM’s subtropical dryness. The GCM also exhibits excessive low-level convective heating, symptomatic of its weak downdrafts; this too is a common feature of other models (Xie et al. 2002). Heating from shallow convection and low stratiform cloud further enhance the overall lower-level heating peak. In the ITCZ and SPCZ ascent regions, the GCM’s excess mean precipitation may explain its underestimate of CWV. The observed stronger gradient in SRF across the Pacific during El Niño in PR data has been suggested by Schumacher et al. (2004) to affect the large-scale dynamical response to El Niño, due to the associated horizontal variability of the diabatic heating vertical structure. We have shown, however, that the GCM’s ENSO-induced SRF behavior is opposite that of PR. This implies that the GCM’s inability to simulate SRF changes during ENSO may be a factor in its weaker remote atmospheric responses to ENSO.

To further investigate the link between ENSO changes in the tropical hydrological cycle and the Hadley and Walker cells, we focus on several regions in the descending and ascending branches of the tropical circulation: 1) The central–east equatorial Pacific (CEEP) region (5°–5°N, 180°–260°E), where the positive ENSO precipitation anomaly is centered; 2) the Maritime Continent (MC) region (15°S–15°N, 90°–150°E), which together with the CEEP region captures the shift of the Walker circulation during El Niño; and 3) the subtropical region in the Northern Hemisphere (STNH; 15°–30°N, 120°–240°E), which covers the descending branch of the Hadley circulation in boreal winter, when ENSO peaks. We focus our discussion on DJF 1998, the peak of the strongest El Niño in the TRMM record.

Horizontal temperature gradients are weak and temperature tendencies small on monthly time scales in the Tropics. The thermodynamic equation can thus to first order be simplified to a balance between diabatic heating/cooling (Q) and adiabatic cooling/warming (ωθ/∂p, where ω is vertical velocity, θ is potential temperature, p is pressure, and ∂θ/∂p is static stability). Perturbations in the diabatic heating (Q′) can be balanced by two contributions, assuming that second-order perturbation terms are small: one from the change in circulation ω′(∂θ/∂p), another from the change in stability ω(∂θ′/∂p), that is,
i1520-0442-20-14-3580-e1
where overbars indicate the climatological value. We estimate each term in (1) for the three regions to understand the GCM’s inability to reproduce the ENSO response.

a. Walker circulation

We first consider the mean and DJF 1998 anomaly behavior of the GCM in the MC and CEEP regions, which constitute the primary upwelling branch of the equatorial Walker cell during normal and ENSO conditions, respectively. Figures 12 and 13 show observed and simulated vertical profiles of latent heating, vertical velocity, static stability, and the two contributions to the adiabatic heating/cooling anomaly in (1) for each region, respectively. Tables 1 and 2 summarize the precipitation and precipitation-type characteristics of the model relative to that observed for each region. (Reanalysis fields in these figures are not totally constrained by observations, but we assume their gross features on seasonal or greater time scales to have some information content.)

Although the GCM overestimates climatological precipitation in the MC relative to both TRMM products (Table 1), its latent heating peaks at lower levels (Fig. 12a) due to its small SRF (23%, versus 43% in PR; Table 1), weak downdraft, and low-level shallow convective and stratiform cloud. As a result, the model’s mean vertical velocity is slightly greater than that in the NCEP–NCAR reanalysis at low levels but considerably weaker in the middle and upper troposphere (Fig. 12b). For the same reason, the temperature profile is too unstable in the middle and upper troposphere throughout the equatorial region (Figs. 12c and 13c). In the CEEP, mean subsidence in the model is too weak at all levels (Fig. 13b). These results are broadly consistent with those of Schumacher et al. (2004, their Fig. 6). The GCM ENSO latent heating anomaly in the CEEP is also centered at low levels, while the TRMM heating anomaly is concentrated at upper levels (Fig. 13a). In fact the GCM SRF anomaly is negative while that observed by PR is positive (Table 2). The remote ENSO subsidence response in the MC is too weak (Fig. 12b) in the model. The vertical velocity anomaly is the primary balance for the diabatic heating anomaly in the Pacific both in the model and observations (Figs. 12d and 13d). Thus, the deficiencies in the model’s remote tropospheric temperature response (Fig. 10c) can be traced to the weakness of its remote subsidence response to ENSO SST perturbations.

This may be understood in the framework of the Gill (1980) model. The remote response depends on the propagation of Kelvin and Rossby waves away from the anomalous heat source. The propagation speed increases with increasing static stability. Since the GCM’s troposphere is too unstable due to weak stratiform heating at upper levels and weak downdraft cooling at lower levels, its wave responses propagate too slowly and thus weaken by dissipation over shorter distances. The resulting weakness of the GCM’s Walker circulation may also explain the overly zonal orientation of its mean and anomalous precipitation fields in the MC and the presence of a primarily north–south rather than west–east dipole ENSO pattern in the IO (Fig. 4).

Another possibility is that the model dissipation is too strong, especially given its use of a cumulus momentum mixing parameterization that neglects cloud-scale pressure gradients. Clarke and Kim (2005) emphasize the role of weak damping in producing a zonally symmetric TT response to ENSO. We performed a sensitivity test in which momentum mixing was suppressed in the convection scheme. The resulting ENSO anomalies are shown in Fig. 14. Eliminating momentum mixing has a beneficial effect on several fields: Spurious positive (negative) ENSO CWV anomalies over the MC (western IO) present in the control simulation are now eliminated. The TT field now shows more of the Gill-like pattern seen in the reanalysis, and cold anomalies are now present over east Asia and Australia. However, negative MC precipitation anomalies are still not produced, and the magnitude of the CEEP ENSO response is still too weak; these features are more likely tied to the deficient upper-level heating.

b. Hadley circulation

The STNH over the Pacific Ocean is the location of the descending branch of the Hadley cell in boreal winter. The excess mean precipitation in the GCM throughout the equatorial region (Tables 1 and 2) implies a meridional gradient of diabatic heating that is too large, and thus a Hadley cell that is too strong, resulting in excess subtropical subsidence (Fig. 15; Table 3). This explains why the GCM underestimates the mean CWV and overestimates the mean OLR in this region (Table 3). On the other hand, the GCM’s lack of stratiform heating and negative SRF ENSO anomaly in the CEEP create a weak subtropical Hadley cell response (Fig. 15), especially in the middle and upper troposphere, so it is no surprise that the model’s CWV and OLR ENSO anomalies are both weaker than observed (Table 3).

In summary, the GISS GCM’s inability to simulate the ENSO remote atmospheric response is mainly a consequence of its underestimates of stratiform rain fraction and downdraft mass flux, and its excessive heating from shallow convective and low stratiform clouds, which result in diabatic heating concentrated in the lower troposphere and a weak upper-tropospheric gradient of anomalous heating. This has potential consequences for the GCM’s estimates of cloud and water vapor feedback in a long-term climate change. Tropical convective cloud feedback depends on changes in the balance between vertical motions, ice production, and sedimentation; the results of this study call into question the model’s ability to simulate these balances. Furthermore, water vapor feedback depends on changes in subtropical Hadley cell subsidence, since it is in these dry regions that OLR is most sensitive to changes in water vapor. The GCM’s weak subtropical ENSO drying anomaly does not necessarily imply that its water vapor feedback is overestimated, since the sense of the Hadley cell change may differ in response to external forcing; it does, however, suggest that the magnitude of water vapor feedback should be considered somewhat more uncertain than might be concluded from the good agreement among current climate models, since most models underestimate SRF.

Together with evidence from the differences in other related fields, we are led to speculate that the lack of a mesoscale updraft, which causes the underestimate of SRF, the weakness of the convective downdraft in the GCM, and the lower-level heating from shallow convection and stratiform cloud, are mainly responsible for its weaker ENSO atmospheric response. For example, the differences between simulated and observed ENSO SRF and ALPH anomalies indicate that lack of a mesoscale updraft limits the GCM’s ability to have anvils respond to ENSO independently of the parent convection. Mesoscale updraft parameterizations are still rare in GCMs (see Donner et al. 2001 for an exception), so this might be a fruitful area for model improvement across the full spectrum of GCMs. No GCM yet produces a reasonable downdraft mass flux (Xie et al. 2002), and some GCMs still have no downdraft parameterization at all; this should also be a priority for model development. In our GCM these weaknesses combine to give an unrealistic heating profile. In fact, it is possible to get a reasonable heating profile without one or both of these parameterization features (Tiedtke 1989; Gregory and Rowntree 1990; Hack 1994; Donner et al. 2001). However, this must be the result of unrealistic updraft mass fluxes or cumulus microphysics, or compensations in other parts of the model, as evidenced by the difficulty these models have in simultaneously getting the cumulus heating and moistening–drying profiles correct. The excess low cloud in the GCM is probably symptomatic of deficiencies in a boundary layer scheme based on dry conserved variables. We also showed that weakening the dissipation associated with cumulus momentum mixing improves some aspects of the remote ENSO response in the simulation. Parameterizations that include the effect of cloud-scale pressure gradients on cumulus momentum transport are available (e.g., Gregory et al. 1997) and should be tested as well. Some of the problems seen in the GISS GCM, for example, its overly zonal response to ENSO (Fig. 4), appear to be a feature of many current GCMs (Dai 2006).

Finally, our comparisons of the various satellite precipitation products have implications both for the use of current products and planning for future missions. Our study makes clear that convective–stratiform partitioning algorithms based on passive microwave observations, while capable of producing reasonable mean geographical distributions, are not yet reliable for studying variability. The planned Global Precipitation Mission (GPM) will rely on a constellation of satellites to provide global coverage on synoptic time scales, but only the core satellite will carry a radar. If GPM is to realize its potential for monitoring convective systems over their life cycle, as the mix of convective and stratiform rainfall changes, further development of passive microwave algorithms, using the TRMM PR as a training device, will be necessary.

Acknowledgments

This research was supported by the NASA Precipitation Measurement Missions and EOS Terra-Aqua Data Analysis Programs. The authors thank Ye Cheng, Jeffrey Jonas, William Kovari, Reto Ruedy, Gary Russell, Mao-Sung Yao, and Yuanchong Zhang for their help in the modeling and data analysis aspects of this work. We also thank two reviewers for helpful comments that improved the manuscript. The TRMM standard products were obtained through the Goddard Earth Sciences Data and Information Service Center’s Distributed Active Archive Center. The other TRMM products were obtained from Remote Sensing Systems. The AMSR-E data were obtained from the National Snow and Ice Data Center. The NCEP–NCAR reanalysis data are distributed by the NOAA Climate Diagnostics Center. The International Satellite Cloud Climatology Project radiative flux dataset (ISCCP-FD) provided the OLR retrievals used in this study.

REFERENCES

  • Adler, R., T. Wilheit Jr., C. Kummerow, and R. Ferraro, cited. 2004: AMSR-E/Aqua monthly L3 5x5 deg rainfall accumulations V001. [Available online at http://www.nsidc.org/data/ae_rngd.html.].

  • Awaka, J., T. Iguchi, H. Kumagai, and K. Okamoto, 1997: Rain type classification algorithm for the TRMM precipitation radar. Proc. IEEE 1997 Int. Geoscience and Remote Sensing Symp., Singapore, IEEE, 1633–1635.

  • Bates, J. J., D. L. Jackson, F-M. Bréon, and Z. D. Bergen, 2001: Variability of tropical upper tropospheric humidity 1979–1998. J. Geophys. Res., 106 , 3227132282.

    • Search Google Scholar
    • Export Citation
  • Berg, W., C. Kummerow, and C. A. Morales, 2002: Differences between east and west Pacific rainfall systems. J. Climate, 15 , 36593672.

    • Search Google Scholar
    • Export Citation
  • Cess, R. D., M. Zhang, P-H. Wang, and B. A. Wielicki, 2001: Cloud structure anomalies over the tropical Pacific during the 1997/98 El Niño. Geophys. Res. Lett., 28 , 45474550.

    • Search Google Scholar
    • Export Citation
  • Chen, J., 2005: Understanding the observed tropical and mid-latitude radiative energy budget in the context of long-term climate variations. Ph.D. thesis, Columbia University, 197 pp.

  • 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 , 26162631.

    • Search Google Scholar
    • Export Citation
  • Chiang, J. C. H., and B. R. Lintner, 2005: Mechanisms of remote tropical surface warming during El Niño. J. Climate, 18 , 41304149.

  • Clarke, A. J., and K-Y. Kim, 2005: On weak zonally symmetric ENSO atmospheric heating and the strong zonally symmetric ENSO air temperature response. J. Atmos. Sci., 62 , 20122022.

    • Search Google Scholar
    • Export Citation
  • Curtis, S., and R. Adler, 2000: ENSO indices based on patterns of satellite-derived precipitation. J. Climate, 13 , 27862793.

  • Dai, A., 2006: Precipitation characteristics in eighteen coupled climate models. J. Climate, 19 , 46054630.

  • Dai, A., I. Y. Fung, and A. D. Del Genio, 1997: Surface observed global land precipitation variations during 1900–88. J. Climate, 10 , 29432962.

    • Search Google Scholar
    • Export Citation
  • Del Genio, A. D., 2002: The dust settles on water vapor feedback. Science, 296 , 665666.

  • Del Genio, A. D., and M. S. Yao, 1993: Efficient cumulus parameterization for long-term climate studies: The GISS scheme. The Representation of Cumulus Convection in Numerical Models, Meteor. Monogr., No. 46, Amer. Meteor. Soc., 181–184.

  • Del Genio, A. D., M-S. Yao, W. Kovari, and K. K-W. Lo, 1996: A prognostic cloud water parameterization for global climate models. J. Climate, 9 , 270304.

    • Search Google Scholar
    • Export Citation
  • Del Genio, A. D., W. Kovari, M-S. Yao, and J. Jonas, 2005: Cumulus microphysics and climate sensitivity. J. Climate, 18 , 23762387.

  • De Moss, J. D., and K. P. Bowman, 2007: Changes in TRMM retrievals due to the orbit boost estimated from rain gauge data. J. Atmos. Oceanic Technol., in press.

    • Search Google Scholar
    • Export Citation
  • Donner, L. J., C. J. Seman, R. S. Hemler, and S. Fan, 2001: A cumulus parameterization including mass fluxes, convective vertical velocities, and mesoscale effects: Thermodynamic and hydrological aspects in a general circulation model. J. Climate, 14 , 34443463.

    • Search Google Scholar
    • Export Citation
  • Fu, C., H. F. Diaz, and J. O. Fletcher, 1986: Characteristics of the response of sea surface temperature in the central Pacific associated with warm episodes of the Southern Oscillation. Mon. Wea. Rev., 114 , 17161739.

    • Search Google Scholar
    • Export Citation
  • Gill, A. E., 1980: Some simple solutions for the heat-induced tropical circulation. Quart. J. Roy. Meteor. Soc., 106 , 447462.

  • Gregory, D., and P. R. Rowntree, 1990: A mass flux convection scheme with representation of cloud ensemble characteristics and stability-dependent closure. Mon. Wea. Rev., 118 , 14831506.

    • Search Google Scholar
    • Export Citation
  • Gregory, D., R. Kershaw, and P. M. Inness, 1997: Parameterization of momentum transport by convection. II: Tests in single-column and general circulation models. Quart. J. Roy. Meteor. Soc., 123 , 11531183.

    • Search Google Scholar
    • Export Citation
  • Hack, J. J., 1994: Parameterization of moist convection in the National Center for Atmospheric Research community climate model (CCM2). J. Geophys. Res., 99 , 55415568.

    • Search Google Scholar
    • Export Citation
  • Hansen, J., and Coauthors, 2002: Climate forcings in Goddard Institute for Space Studies SI2000 simulations. J. Geophys. Res., 107 .4347, doi:10.1029/2001JD001143.

    • Search Google Scholar
    • Export Citation
  • Houze Jr., R. A., 1982: Cloud clusters and large-scale vertical motions in the tropics. J. Meteor. Soc. Japan, 60 , 396409.

  • Houze Jr., R. A., 1997: Stratiform precipitation in regions of convection: A meteorological paradox? Bull. Amer. Meteor. Soc., 78 , 21792196.

    • Search Google Scholar
    • Export Citation
  • Huffman, G. J., and Coauthors, 1997: The Global Precipitation Climatology Project (1997) combined precipitation dataset. Bull. Amer. Meteor. Soc., 78 , 520.

    • Search Google Scholar
    • Export Citation
  • Iguchi, T., T. Kozu, R. Meneghini, J. Awaka, and K. Okamoto, 2000: Rain-profiling algorithm for the TRMM precipitation radar. J. Appl. Meteor., 39 , 20382052.

    • Search Google Scholar
    • Export Citation
  • Kalnay, E., and Coauthors, 1996: The NCEP/NCAR 40-Year Reanalysis Project. Bull. Amer. Meteor. Soc., 77 , 437471.

  • Kelly, P. M., and P. D. Jones, 1996: Removal of the El Niño–Southern Oscillation signal from the gridded surface air temperature data set. J. Geophys. Res., 101 , 1901319022.

    • Search Google Scholar
    • Export Citation
  • 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 , 917932.

    • Search Google Scholar
    • Export Citation
  • Kug, J-S., and I-S. Kang, 2006: Interactive feedback between ENSO and the Indian Ocean. J. Climate, 19 , 17841801.

  • Kummerow, C., and Coauthors, 2000: The status of the Tropical Rainfall Measuring Mission (TRMM) after two years in orbit. J. Appl. Meteor., 39 , 19651982.

    • Search Google Scholar
    • Export Citation
  • Kummerow, C., and Coauthors, 2001: The evolution of the Goddard Profiling Algorithm (GPROF) for rainfall estimation from passive microwave sensors. J. Appl. Meteor., 40 , 18011820.

    • Search Google Scholar
    • Export Citation
  • Lanzante, J. R., 1996: Lag relationships involving tropical sea surface temperatures. J. Climate, 9 , 25682578.

  • Lu, R., B. Dong, R. D. Cess, and G. L. Potter, 2004: The 1997/98 El Niño: A test for climate models. Geophys. Res. Lett., 31 .L12216, doi:10.1029/2004GL019956.

    • Search Google Scholar
    • Export Citation
  • Olson, W. S., C. D. Kummerow, Y. Hong, and W-K. Tao, 1999: Atmospheric latent heating distributions in the Tropics derived from satellite passive microwave radiometer measurements. J. Appl. Meteor., 38 , 633664.

    • Search Google Scholar
    • Export Citation
  • Olson, W. S., Y. Hong, C. D. Kummerow, and J. Turk, 2001: A texture-polarization method for estimating convective–stratiform precipitation area coverage from passive microwave radiometer data. J. Appl. Meteor., 40 , 15771591.

    • Search Google Scholar
    • Export Citation
  • Olson, W. S., and Coauthors, 2006: Precipitation and latent heating distributions from satellite passive microwave radiometry. Part I: Improved method and uncertainties. J. Appl. Meteor. Climatol., 45 , 702720.

    • Search Google Scholar
    • Export Citation
  • Rayner, N. A., D. E. Parker, E. B. Horton, C. K. Folland, L. V. Alexander, D. P. Rowell, E. C. Kent, and A. Kaplan, 2003: Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century. J. Geophys. Res., 108 .4407, doi:10.1029/2002JD002670.

    • Search Google Scholar
    • Export Citation
  • Robertson, F. R., R. W. Spencer, and D. E. Fitzjarrald, 2001: A new satellite deep convective ice index for tropical climate monitoring: Possible implications for existing oceanic precipitation datasets. Geophys. Res. Lett., 28 , 251254.

    • Search Google Scholar
    • Export Citation
  • Robertson, F. R., D. E. Fitzjarrald, and C. D. Kummerow, 2003: Effects of uncertainty in TRMM precipitation radar path integrated attenuation on interannual variations of tropical oceanic rainfall. Geophys. Res. Lett., 30 .1180, doi:10.1029/2002GL016416.

    • Search Google Scholar
    • Export Citation
  • Robock, A., and J. Mao, 1995: The volcanic signal in surface temperature observations. J. Climate, 8 , 10861103.

  • Saji, N. H., B. N. Goswami, P. N. Vinayachandran, and T. Yamagata, 1999: A dipole mode in the tropical Indian Ocean. Nature, 401 , 360363.

    • Search Google Scholar
    • Export Citation
  • Santer, B. D., and Coauthors, 2001: Accounting for the effects of volcanoes and ENSO in comparisons of modeled and observed temperature trends. J. Geophys. Res., 106 , 2803328060.

    • Search Google Scholar
    • Export Citation
  • Schmidt, G. A., and Coauthors, 2006: Present-day atmospheric simulations using GISS ModelE: Comparison to in situ, satellite, and reanalysis data. J. Climate, 19 , 153192.

    • Search Google Scholar
    • Export Citation
  • Schumacher, C., and R. A. Houze Jr., 2003: Stratiform rain in the Tropics as seen by the TRMM precipitation radar. J. Climate, 16 , 17391756.

    • Search Google Scholar
    • Export Citation
  • Schumacher, C., R. A. Houze Jr., and I. Kraucunas, 2004: The tropical dynamical response to latent heating estimates derived from the TRMM Precipitation Radar. J. Atmos. Sci., 61 , 13411358.

    • Search Google Scholar
    • Export Citation
  • Shimizu, S., N. Takahashi, T. Iguchi, J. Awaka, T. Kozu, R. Meneghini, and K. Okamoto, 2003: Validation analyses after the altitude change of TRMM. Proc. SPIE, 4894 , 8391.

    • Search Google Scholar
    • Export Citation
  • Soden, B. J., 2000: The sensitivity of the tropical hydrological cycle to ENSO. J. Climate, 13 , 538549.

  • Tao, W. K., S. Lang, J. Simpson, and R. Adler, 1993: Retrieval algorithms for estimating the vertical profiles of latent heat release: Their applications for TRMM. J. Meteor. Soc. Japan, 71 , 685700.

    • Search Google Scholar
    • Export Citation
  • Tao, W. K., and Coauthors, 2001: Retrieved vertical profiles of latent heat release using TRMM rainfall products for February 1998. J. Appl. Meteor., 40 , 957982.

    • Search Google Scholar
    • Export Citation
  • Tiedtke, M., 1989: A comprehensive mass flux scheme for cumulus parameterization in large-scale models. Mon. Wea. Rev., 117 , 17791800.

    • Search Google Scholar
    • Export Citation
  • Trenberth, K. E., and J. W. Hurrell, 1994: Decadal atmosphere–ocean variations in the Pacific. Climate Dyn., 9 , 303319.

  • Trenberth, K. E., J. M. Caron, D. P. Stepaniak, and S. Worley, 2002: Evolution of El Niño–Southern Oscillation and global atmospheric surface temperatures. J. Geophys. Res., 107 .4065, doi:10.1029/2000JD000298.

    • Search Google Scholar
    • Export Citation
  • Wentz, F. J., and T. Meissner, 2000: AMSR Ocean Algorithm Version 2. RSS Tech. Rep. 121599A, Remote Sensing Systems, Santa Rosa, CA, 59 pp.

  • Wilheit, T., C. Kummerow, and R. Ferraro, 2003: Rainfall algorithms for AMSR-E. IEEE Trans. Geosci. Remote Sens., 41 , 204214.

  • Xie, P., and P. A. Arkin, 1998: Global monthly precipitation estimates from satellite-observed outgoing longwave radiation. J. Climate, 11 , 137164.

    • Search Google Scholar
    • Export Citation
  • Xie, S., and Coauthors, 2002: Intercomparison and evaluation of cumulus parameterizations under summertime midlatitude continental conditions. Quart. J. Roy. Meteor. Soc., 128 , 10951135.

    • Search Google Scholar
    • Export Citation
  • 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 , 17191736.

    • Search Google Scholar
    • Export Citation
  • Zhang, Y., W. B. Rossow, A. A. Lacis, V. Oinas, and M. I. Mishchenko, 2004: Calculation of radiative fluxes from the surface to top of atmosphere based on ISCCP and other global data sets: Refinements of the radiative transfer model and the input data. J. Geophys. Res., 109 .D19105, doi:10.10129/2003JD004457.

    • Search Google Scholar
    • Export Citation

APPENDIX

ENSO Extraction Procedure

We first calculate the extreme correlation (of either sign) at the least lag (ECLL) between the Niño-3.4 index (N34) and the anomalies of a given climate variable X at each grid box. The choice of N34 is based on the study of Trenberth et al. (2002), which suggested that N34 is best suited to characterize ENSO lag–lead relationships. The N34 used in this study is derived from the same HadISST1 dataset that is used to force the GCM simulation (available online at http://www.cdc.noaa.gov/Pressure/Timeseries/Nino34/). Chen (2005) shows that at equatorial and subtropical latitudes in both hemispheres, the N34–SST correlation pattern as a function of lag at most longitudes has a single well-defined peak. This makes ECLL useful for defining a remote ENSO signal that accounts for the fact that the lag can vary geographically depending on the sequence of processes that remotely influence the SST (Klein et al. 1999).

The next step is to calculate the standard deviations of N34 (σN34) and of X (σX) at each grid box. Then the ENSO component of variability X′ is given as
i1520-0442-20-14-3580-ea1
where index i refers to the grid box, m the time, and j the lag in months. According to (A1) the ENSO component of the variability in parameter X at a given lag relative to N34 simply depends on the magnitude of variability of X relative to N34 and the extent to which these variabilities are correlated. The biggest limitation of the method is its dependence on a linear cross-correlation between N34 and ENSO variations in other climate fields. Thus it captures only the stable, linear, and symmetric aspects of the ENSO-induced signal.

Chen (2005) tests the ENSO extraction procedure in two ways. First, he performs empirical orthogonal function (EOF) analysis of several independent multidecadal global surface temperature datasets and correlates the resulting principal components with N34 at lags from −12 to +12 months to define the leading modes that capture the ENSO signal. The extraction method defined by (A1) is then applied and the result subtracted from the original time series. The EOF analysis is repeated on the resulting modified datasets and N34 correlated with the six leading principal components at the same set of lags. He finds that none of the six leading modes is now correlated with N34 at any lag at the 99% confidence level. The leading modes in the adjusted time series instead portray a century-long upward temperature trend and a decadal variability mode that resembles the Pacific decadal oscillation (Trenberth and Hurrell 1994).

Second, Chen compares the EOFs associated with ENSO and the decadal variability mode that emerges when ENSO is removed in the surface temperature field and a variety of NCEP–NCAR reanalysis fields (horizontal and vertical velocity, Walker cell overturning, air temperature, and specific humidity). The ENSO anomaly fields are found to have characteristic, well-documented spatial patterns and differ noticeably from the decadal patterns. The ENSO signal, for example, being of opposite sign to the decadal mode over the Maritime Continent and much of eastern North America, is considerably stronger and somewhat east-shifted relative to the decadal mode in the equatorial east-central Pacific, stronger also over South America, the tropical Atlantic, and southeastern United States, but weaker in the subtropical and north extratropical Pacific.

Fig. 1.
Fig. 1.

Seasonal mean ENSO anomaly evolution in the RSS–TMI derived SST field. Seasons are DJF, March–May (MAM), June–August (JJA), and September–November (SON).

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 2.
Fig. 2.

The extreme correlation at least lag with (top) Niño-3.4 and (bottom) the corresponding lag relationship maps for TMI-derived SST. Only those locations above the 95% significance level in the lag map are shown.

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 3.
Fig. 3.

Dec 1997–Jun 2005 mean annual precipitation from (a) TMI, (b) PR, (c) the GCM, and (d) zonal means of all three.

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 4.
Fig. 4.

Seasonal ENSO anomalies in (a) TMI and (b) GCM precipitation.

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 4.
Fig. 4.

(Continued)

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 5.
Fig. 5.

ENSO monthly precipitation anomaly scatterplots for (a),(b) PR vs (c),(d) TMI and (e),(f) AMSR-E vs TMI. (left) El Niño months and (right) La Niña months as defined in the text. The solid line is the least squares fit; S stands for the slope of the linear fit, and R is the correlation coefficient.

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 6.
Fig. 6.

The extreme correlation at least lag of precipitation with Niño-3.4 and the corresponding lag relationship maps for (a) PR, (b) TMI, and (c) the GCM. Only those locations above the 95% significance level in the lag map are shown.

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 7.
Fig. 7.

(left) Mean and (right) ENSO anomaly in stratiform rainfall fraction in DJF 1998 from (a),(d) PR, (b),(e) TMI, and (c),(f) the GCM.

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 8.
Fig. 8.

(left) Mean and (right) ENSO anomaly for the altitude of peak latent heating (APLH) for (a),(c) TRMM CSH and (b),(d) the GCM.

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 9.
Fig. 9.

GCM mean DJF latent heating profiles for (a) the Maritime Continent (15°S–15°N, 90°–150°E) and (b) the central-eastern equatorial Pacific (5°S–5°N, 180°–260°E). The total heating (solid line) is decomposed into shallow convective (dashed–dotted), deep convective (dotted), and stratiform cloud (long dashed) contributions.

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 10.
Fig. 10.

1998 DJF ENSO anomaly in (a) column water vapor, (b) outgoing longwave radiation, and (c) middle-tropospheric temperature in (left) the observations and (right) the GCM.

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 11.
Fig. 11.

Scatterplot of ENSO anomalies in CWV vs 500-mb omega over the subtropics in both hemispheres (15°–30°S, 15°–30°N) in (top) observations/reanalysis and (bottom) the GISS GCM. The solid line is the least squares fit. Here, S ≡ slope of the linear fit, and R ≡ correlation coefficient.

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 12.
Fig. 12.

Vertical profiles of each element in (1) for climatology and for the DJF 1998 difference with respect to climatology for the Maritime Continent region (15°S–15°N, 90°–150°E) including (a) latent heating contribution to Q and Q′, (b) ω and ω′, (c) (∂θ/∂p) and (∂θ′/∂p), and (d) − ω(∂θ′/∂p) (term 1) and − ω′(∂θ/∂p) (term 2). The mean latent heating rate is scaled by a factor of 0.1 in order to fit the anomaly in the same plot. In (a)–(c), the solid (mean) and dashed (perturbation) lines are for the GCM, and dotted (mean) and dashed–dotted (perturbation) lines are for NCEP–NCAR except for the diabatic heating term, which is from TRMM CSH. In (d), all curves are for the perturbation, with the solid and dotted lines indicating term 1 and the dashed and dashed–dotted lines indicating term 2.

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 13.
Fig. 13.

Same as in Fig. 12, but for the central-eastern equatorial Pacific region (5°S–5°N, 180°–260°E).

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 14.
Fig. 14.

GCM ENSO anomalies in (a) precipitation, (b) CWV, (c), OLR, and (d) TT for the simulation without cumulus momentum mixing.

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Fig. 15.
Fig. 15.

Omega profiles as in Fig. 12b, but for the subtropical region in the Northern Hemisphere (15°–30°N, 120°–240°E).

Citation: Journal of Climate 20, 14; 10.1175/JCLI4208.1

Table 1.

PREC and SRF means and DJF 1998 anomalies over the Maritime Continent.

Table 1.
Table 2.

Same as in Table 1, but for the central–eastern equatorial Pacific region.

Table 2.
Table 3.

Mean and DJF 1998 anomaly for the subtropical Pacific region in the Northern Hemisphere in omega at 500 mb, OLR, and CWV.

Table 3.
Save
  • Adler, R., T. Wilheit Jr., C. Kummerow, and R. Ferraro, cited. 2004: AMSR-E/Aqua monthly L3 5x5 deg rainfall accumulations V001. [Available online at http://www.nsidc.org/data/ae_rngd.html.].

  • Awaka, J., T. Iguchi, H. Kumagai, and K. Okamoto, 1997: Rain type classification algorithm for the TRMM precipitation radar. Proc. IEEE 1997 Int. Geoscience and Remote Sensing Symp., Singapore, IEEE, 1633–1635.

  • Bates, J. J., D. L. Jackson, F-M. Bréon, and Z. D. Bergen, 2001: Variability of tropical upper tropospheric humidity 1979–1998. J. Geophys. Res., 106 , 3227132282.

    • Search Google Scholar
    • Export Citation
  • Berg, W., C. Kummerow, and C. A. Morales, 2002: Differences between east and west Pacific rainfall systems. J. Climate, 15 , 36593672.

    • Search Google Scholar
    • Export Citation
  • Cess, R. D., M. Zhang, P-H. Wang, and B. A. Wielicki, 2001: Cloud structure anomalies over the tropical Pacific during the 1997/98 El Niño. Geophys. Res. Lett., 28 , 45474550.

    • Search Google Scholar
    • Export Citation
  • Chen, J., 2005: Understanding the observed tropical and mid-latitude radiative energy budget in the context of long-term climate variations. Ph.D. thesis, Columbia University, 197 pp.

  • 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 , 26162631.

    • Search Google Scholar
    • Export Citation
  • Chiang, J. C. H., and B. R. Lintner, 2005: Mechanisms of remote tropical surface warming during El Niño. J. Climate, 18 , 41304149.

  • Clarke, A. J., and K-Y. Kim, 2005: On weak zonally symmetric ENSO atmospheric heating and the strong zonally symmetric ENSO air temperature response. J. Atmos. Sci., 62 , 20122022.

    • Search Google Scholar
    • Export Citation
  • Curtis, S., and R. Adler, 2000: ENSO indices based on patterns of satellite-derived precipitation. J. Climate, 13 , 27862793.

  • Dai, A., 2006: Precipitation characteristics in eighteen coupled climate models. J. Climate, 19 , 46054630.

  • Dai, A., I. Y. Fung, and A. D. Del Genio, 1997: Surface observed global land precipitation variations during 1900–88. J. Climate, 10 , 29432962.

    • Search Google Scholar
    • Export Citation
  • Del Genio, A. D., 2002: The dust settles on water vapor feedback. Science, 296 , 665666.

  • Del Genio, A. D., and M. S. Yao, 1993: Efficient cumulus parameterization for long-term climate studies: The GISS scheme. The Representation of Cumulus Convection in Numerical Models, Meteor. Monogr., No. 46, Amer. Meteor. Soc., 181–184.

  • Del Genio, A. D., M-S. Yao, W. Kovari, and K. K-W. Lo, 1996: A prognostic cloud water parameterization for global climate models. J. Climate, 9 , 270304.

    • Search Google Scholar
    • Export Citation
  • Del Genio, A. D., W. Kovari, M-S. Yao, and J. Jonas, 2005: Cumulus microphysics and climate sensitivity. J. Climate, 18 , 23762387.

  • De Moss, J. D., and K. P. Bowman, 2007: Changes in TRMM retrievals due to the orbit boost estimated from rain gauge data. J. Atmos. Oceanic Technol., in press.

    • Search Google Scholar
    • Export Citation
  • Donner, L. J., C. J. Seman, R. S. Hemler, and S. Fan, 2001: A cumulus parameterization including mass fluxes, convective vertical velocities, and mesoscale effects: Thermodynamic and hydrological aspects in a general circulation model. J. Climate, 14 , 34443463.

    • Search Google Scholar
    • Export Citation
  • Fu, C., H. F. Diaz, and J. O. Fletcher, 1986: Characteristics of the response of sea surface temperature in the central Pacific associated with warm episodes of the Southern Oscillation. Mon. Wea. Rev., 114 , 17161739.

    • Search Google Scholar
    • Export Citation
  • Gill, A. E., 1980: Some simple solutions for the heat-induced tropical circulation. Quart. J. Roy. Meteor. Soc., 106 , 447462.

  • Gregory, D., and P. R. Rowntree, 1990: A mass flux convection scheme with representation of cloud ensemble characteristics and stability-dependent closure. Mon. Wea. Rev., 118 , 14831506.

    • Search Google Scholar
    • Export Citation
  • Gregory, D., R. Kershaw, and P. M. Inness, 1997: Parameterization of momentum transport by convection. II: Tests in single-column and general circulation models. Quart. J. Roy. Meteor. Soc., 123 , 11531183.

    • Search Google Scholar
    • Export Citation
  • Hack, J. J., 1994: Parameterization of moist convection in the National Center for Atmospheric Research community climate model (CCM2). J. Geophys. Res., 99 , 55415568.

    • Search Google Scholar
    • Export Citation
  • Hansen, J., and Coauthors, 2002: Climate forcings in Goddard Institute for Space Studies SI2000 simulations. J. Geophys. Res., 107 .4347, doi:10.1029/2001JD001143.

    • Search Google Scholar
    • Export Citation
  • Houze Jr., R. A., 1982: Cloud clusters and large-scale vertical motions in the tropics. J. Meteor. Soc. Japan, 60 , 396409.

  • Houze Jr., R. A., 1997: Stratiform precipitation in regions of convection: A meteorological paradox? Bull. Amer. Meteor. Soc., 78 , 21792196.

    • Search Google Scholar
    • Export Citation
  • Huffman, G. J., and Coauthors, 1997: The Global Precipitation Climatology Project (1997) combined precipitation dataset. Bull. Amer. Meteor. Soc., 78 , 520.

    • Search Google Scholar
    • Export Citation
  • Iguchi, T., T. Kozu, R. Meneghini, J. Awaka, and K. Okamoto, 2000: Rain-profiling algorithm for the TRMM precipitation radar. J. Appl. Meteor., 39 , 20382052.

    • Search Google Scholar
    • Export Citation
  • Kalnay, E., and Coauthors, 1996: The NCEP/NCAR 40-Year Reanalysis Project. Bull. Amer. Meteor. Soc., 77 , 437471.

  • Kelly, P. M., and P. D. Jones, 1996: Removal of the El Niño–Southern Oscillation signal from the gridded surface air temperature data set. J. Geophys. Res., 101 , 1901319022.

    • Search Google Scholar
    • Export Citation
  • 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 , 917932.

    • Search Google Scholar
    • Export Citation
  • Kug, J-S., and I-S. Kang, 2006: Interactive feedback between ENSO and the Indian Ocean. J. Climate, 19 , 17841801.

  • Kummerow, C., and Coauthors, 2000: The status of the Tropical Rainfall Measuring Mission (TRMM) after two years in orbit. J. Appl. Meteor., 39 , 19651982.

    • Search Google Scholar
    • Export Citation
  • Kummerow, C., and Coauthors, 2001: The evolution of the Goddard Profiling Algorithm (GPROF) for rainfall estimation from passive microwave sensors. J. Appl. Meteor., 40 , 18011820.

    • Search Google Scholar
    • Export Citation
  • Lanzante, J. R., 1996: Lag relationships involving tropical sea surface temperatures. J. Climate, 9 , 25682578.

  • Lu, R., B. Dong, R. D. Cess, and G. L. Potter, 2004: The 1997/98 El Niño: A test for climate models. Geophys. Res. Lett., 31 .L12216, doi:10.1029/2004GL019956.

    • Search Google Scholar
    • Export Citation
  • Olson, W. S., C. D. Kummerow, Y. Hong, and W-K. Tao, 1999: Atmospheric latent heating distributions in the Tropics derived from satellite passive microwave radiometer measurements. J. Appl. Meteor., 38 , 633664.

    • Search Google Scholar
    • Export Citation
  • Olson, W. S., Y. Hong, C. D. Kummerow, and J. Turk, 2001: A texture-polarization method for estimating convective–stratiform precipitation area coverage from passive microwave radiometer data. J. Appl. Meteor., 40 , 15771591.

    • Search Google Scholar
    • Export Citation
  • Olson, W. S., and Coauthors, 2006: Precipitation and latent heating distributions from satellite passive microwave radiometry. Part I: Improved method and uncertainties. J. Appl. Meteor. Climatol., 45 , 702720.

    • Search Google Scholar
    • Export Citation
  • Rayner, N. A., D. E. Parker, E. B. Horton, C. K. Folland, L. V. Alexander, D. P. Rowell, E. C. Kent, and A. Kaplan, 2003: Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century. J. Geophys. Res., 108 .4407, doi:10.1029/2002JD002670.

    • Search Google Scholar
    • Export Citation
  • Robertson, F. R., R. W. Spencer, and D. E. Fitzjarrald, 2001: A new satellite deep convective ice index for tropical climate monitoring: Possible implications for existing oceanic precipitation datasets. Geophys. Res. Lett., 28 , 251254.

    • Search Google Scholar
    • Export Citation
  • Robertson, F. R., D. E. Fitzjarrald, and C. D. Kummerow, 2003: Effects of uncertainty in TRMM precipitation radar path integrated attenuation on interannual variations of tropical oceanic rainfall. Geophys. Res. Lett., 30 .1180, doi:10.1029/2002GL016416.

    • Search Google Scholar
    • Export Citation
  • Robock, A., and J. Mao, 1995: The volcanic signal in surface temperature observations. J. Climate, 8 , 10861103.

  • Saji, N. H., B. N. Goswami, P. N. Vinayachandran, and T. Yamagata, 1999: A dipole mode in the tropical Indian Ocean. Nature, 401 , 360363.

    • Search Google Scholar
    • Export Citation
  • Santer, B. D., and Coauthors, 2001: Accounting for the effects of volcanoes and ENSO in comparisons of modeled and observed temperature trends. J. Geophys. Res., 106 , 2803328060.

    • Search Google Scholar
    • Export Citation
  • Schmidt, G. A., and Coauthors, 2006: Present-day atmospheric simulations using GISS ModelE: Comparison to in situ, satellite, and reanalysis data. J. Climate, 19 , 153192.

    • Search Google Scholar
    • Export Citation
  • Schumacher, C., and R. A. Houze Jr., 2003: Stratiform rain in the Tropics as seen by the TRMM precipitation radar. J. Climate, 16 , 17391756.

    • Search Google Scholar
    • Export Citation
  • Schumacher, C., R. A. Houze Jr., and I. Kraucunas, 2004: The tropical dynamical response to latent heating estimates derived from the TRMM Precipitation Radar. J. Atmos. Sci., 61 , 13411358.

    • Search Google Scholar
    • Export Citation
  • Shimizu, S., N. Takahashi, T. Iguchi, J. Awaka, T. Kozu, R. Meneghini, and K. Okamoto, 2003: Validation analyses after the altitude change of TRMM. Proc. SPIE, 4894 , 8391.

    • Search Google Scholar
    • Export Citation
  • Soden, B. J., 2000: The sensitivity of the tropical hydrological cycle to ENSO. J. Climate, 13 , 538549.

  • Tao, W. K., S. Lang, J. Simpson, and R. Adler, 1993: Retrieval algorithms for estimating the vertical profiles of latent heat release: Their applications for TRMM. J. Meteor. Soc. Japan, 71 , 685700.

    • Search Google Scholar
    • Export Citation
  • Tao, W. K., and Coauthors, 2001: Retrieved vertical profiles of latent heat release using TRMM rainfall products for February 1998. J. Appl. Meteor., 40 , 957982.