Climate Dynamics (2004) 22: 205–222 DOI 10.1007/s00382-003-0376-7 A. Gershunov Æ R. Roca Coupling of latent heat flux and the greenhouse effect by large-scale tropical /subtropical dynamics diagnosed in a set of observations and model simulations Received: 19 February 2003 / Accepted: 14 October 2003 / Published online: 28 January 2004 Springer-Verlag 2004 Abstract Coupled variability of the greenhouse effect (GH) and latent heat flux (LHF) over the tropical – subtropical oceans is described, summarized and compared in observations and a coupled ocean-atmosphere general circulation model (CGCM). Coupled seasonal and interannual modes account for much of the total variability in both GH and LHF. In both observations and model, seasonal coupled variability is locally 180 out-of-phase throughout the tropics. Moisture is brought into convergent/convective regions from remote source areas located partly in the opposite, non-convective hemisphere. On interannual time scales, the tropical Pacific GH in the ENSO region of largest interannual variance is 180 out of phase with local LHF in observations but in phase in the model. A local source of moisture is thus present in the model on interannual time scales while in observations, moisture is mostly advected from remote source regions. The latent cooling and radiative heating of the surface as manifested in the interplay of LHF and GH is an important determinant of the current climate. Moreover, the hydrodynamic processes involved in the GH–LHF interplay determine in large part the climate response to external perturbations mainly through influencing the water vapor feedback but also through their intimate connection to the hydrological cycle. The diagnostic process proposed here can be performed on other CGCMs. Similarly, it should be repeated using a number of observational latent heat flux datasets to account for the variability in the A. Gershunov (&) Climate Research Division, Scripps Institution of Oceanography, La Jolla, CA 92093-0224, USA, E-mail: sasha@ucsd.edu R. Roca Laboratoire de Météorologie Dynamique, Ecole Polytechnique, 91128 Palaiseau, France different satellite retrievals. A realistic CGCM could be used to further study these coupled dynamics in natural and anthropogenically altered climate conditions. 1 Introduction Water vapor is the most important greenhouse gas in the atmosphere. The radiative properties of water vapor are central to the response of the climate system to perturbations. The water vapor feedback is usually considered as a strong positive feedback on sea surface temperature (e.g., Manabe and Wetherald 1967; Held and Soden 2000). While substantial debate has taken place about the functioning of this feedback in the past decade (e.g. Lindzen 1990; Pierrehumbert 1995; Zhu et al. 2000), the processes at play are still poorly understood. The classical view, inherited from global radiative-convective one dimensional model studies, links the atmospheric water vapor content to atmospheric temperature by simple thermodynamics (the Clausius-Clapeyron law) as well as to global surface evaporation acting as a source of moisture for the atmosphere (Ramanathan 1981). One feature lacking in this classical view concerns the role of large-scale dynamics and water vapor transport from surface sources to atmospheric sinks. In the intertropical belt, evaporation maxima are located over the subtropical regions, which act as sources of moisture for the atmosphere. The low-level Hadley/Walker circulations transport this moisture from the subtropical sources towards the meteorological equator and warm pools (Cornejo-Garrido and Stone 1977; Gershunov et al. 1998) where deep convection injects moisture into the free troposphere. The advected water vapor and liquid water (a product of convection) form regions of strong greenhouse effect in tropical convergence zones while subtropical high pressure regions are cooled by releasing latent heat of evaporation. The non-local nature of the GH–LHF coupling was investigated by Gershunov et al. (1998, hereafter G98), 206 Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale who used a combination of satellite and in situ observations as well as analyzed surface wind fields and lowlevel moisture convergence to show that the clear-sky GH (GHcs) is locally anticorrelated with LHF while strong positive temporal correlations between GHcs and LHF exist in remote regions of surface moisture sources and atmospheric sinks. These remote links are imbedded in seasonally and interannually varying Hadley and Walker circulations. On seasonal time scales, much of the moisture feeding convergence zones in the summer hemisphere is advected across the Equator from the subtropical high-pressure zones of the winter hemisphere. This non-local relationship between GH and LHF shown on climatic time scales throughout the tropics indicates that LHF can not provide either a local source for the greenhouse, or a local control on greenhouse warming. Understanding the energy balance of the tropics involves the consideration of radiative and thermodynamic variables in the context of the entire atmospheric circulation as it varies on all relevant climatic time scales. The mechanisms involved in maintaining the spatial distribution and variability of GH and LHF may influence natural climate variability through their potential to influence the circulations in which they are imbedded. [NB As a basic example, consider the following illustration. GH promotes atmospheric and surface warming in the western Pacific warm pool region, enhancing zonal temperature and pressure gradients which promote the pacific trades enhancing subsidence, divergence, evaporation, and latent surface cooling in the eastern Pacific subtropical highs. This works to further enhance horizontal temperature and pressure gradients. Consequently, water vapor is more efficiently transported across the Pacific where it feeds GH and convection over the warm pool further enhancing the Walker and Hadley circulations (G98).] These same mechanisms play a paramount role in determining the nature of anthropogenic climate change by controlling the water vapor feedback to external climate perturbations, thought to be the largest positive feedback in the climate system (e.g., Manabe and Wetherald 1967; Held and Soden 2000). In the present study, the results of G98 are reinvestigated using a novel dataset and observational period as well as a statistical technique dedicated to the analysis and efficient description of the coupling between LHF and GH. In order to account for the longwave radiative effect of atmospheric moisture in all its phases, total sky GH is considered instead of the clear sky greenhouse effect previously investigated in G98. Furthermore, the coupling is examined in a fully coupled global ocean-atmosphere model (CGCM) to test whether the CGCM can be used to better understand the mechanisms that control the distribution and co-variability of GH and LHF, especially at time scales longer than those covered by the observational record, i.e., decadal variability and climate change experiments. The work is organized as follows. Section 2 introduces the different datasets used to perform the analysis. A brief description of the CGCM simulations is also included. Section 3 provides a broad statistical description of the observations and model output, i.e., means, variances and local temporal relationships between GH and LHF, revealing the necessity for a coupled non-local analysis. In Sect. 4, the seasonal coupling between LHF and GH is assessed via canonical correlation analysis. The behavior of the CGCM is compared to observed reality. Section 5 addresses modeled and observed GH– LHF variability and coupling on interannual time scales. Finally, discussion and conclusions are provided in Sect. 6. 2 Data 2.1 Observations The total sky greenhouse effect is computed using the Reynolds optimum interpolation sea surface temperature (SST) analysis (Reynolds 1988) and total sky outgoing longwave radiation (OLR) obtained from NOAA (Smith et al. 1996). GH is only computed over the oceans, which are assumed to emit like a blackbody. This introduces an uncertainty of less than 1% in the computations (Inamdar and Ramanathan 1998). The OLR is derived from the operational NOAA satellites, which measure narrow band radiation, which is converted into broadband. When compared to the Earth Radiation Budget Experiment (ERBE) data and to the most recent ScaraB-1, -2 and CERES flux, the NOAA product shows a cold bias of 8 W/m2 (e.g., Duvel et al. 2001). While such a data set is not well suited for analyzing long-term climate trends, the seasonal and interannual variability, of interest in the present study, is in good agreement with the other satellite measurements. The NOAA products were preferred to the time-limited ERBE estimates of TOA radiation in order to span the observed LHF more recent time period.The resulting 8 W/m2 GH bias towards higher values does not influence our results on variability and the coupling between GH and LHF, but it must be taken into account when we consider the mean GH field.The period of study spans January 1992 to December 1996. Monthly means are used at a 2.5 · 2.5 resolution. The latent heat flux (LHF) observations are provided by the Hamburg Ocean Atmosphere Parameters and Fluxes from Satellite Data (HOAPS) products (Grassl et al. 2000). It consists of monthly mean estimates of LHF at a 2.5 · 2.5 resolution. To estimate the LHF, a bulk formulation is followed where each term is evaluated individually. The SSTs are retrieved from the AVHRR sensor of the NOAA/NASA Pathfinder Ocean Program (Smith et al. 1996). The wind speed u and the atmospheric specific humidity q are determined from SSM/I satellite measurements while the transfer coefficient is determined empirically (Schulz et al. 1993; Schluessel et al. 1995). Schulz et al. (1997) compared the results to in-situ measurements during TOGA-COARE and indicated that in the tropical conditions of interest, the uncertainty in LHF reaches 15 W/m2 for monthly mean time scale. Despite this bias, the time variability of the satellite product was shown to agree reasonably well with in situ observations. In the present study we focus on the most stable period of the HOAPS climatology spanning January 1992 to December 1996 when a single SMM/I satellite was operational, avoiding issues of intercalibration and long-term inhomogeneities. The three datasets (SST, OLR and LHF) suffer from small estimated biases, which should be borne in mind when direct comparisons are made with the CGCM output. Recent intercomparison between different satellite-based climatologies of LHF indeed indicates that the HOAPS climatology underestimates the mean LHF in the tropics with respect to other available products (Kubota et al. 2003). However as discussed later, the methodology employed here mainly relies upon correlations and hence the analysis of the coupled GH–LHF variability is not corrupted by Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale these systematic biases in the mean fields. A thorough intercomparison of the variability structure in the different available data sets of OLR and LHF should shed more light on the discrepancies between each data set. Such an analysis is outside the scope of the present study. We proceed, bearing in mind the limitations of the present observational data. In addition to the listed thermodynamic variables, we used the 850 hPa re-analyzed winds from the National Center for Environmental Prediction (NCEP) (Kalnay et al. 1996) to document the large-scale low-level flow. 2.2 CGCM The Community Climate System Model (CCSM) is one of the most advanced coupled models currently available. Being the first climate model developed and applied by the scientific community using pooled resources from many institutions, it relies on and benefits from the involvement of the broad climate research community for its advancement. The CCSM includes interactive atmosphere, ocean, land, and sea ice models whose details and coupling are described by Blackmon et al. (2001). One of the long-term goals of the CCSM project is ‘‘to make the model readily available to, and usable by, the climate research community, and to actively engage the community in the ongoing process of model development’’ (Blackmon et al. 2001). The source code with documentation and output from the primary CCSM simulations have both been made available on the CCSM Web site (http://www.ccsm.ucar.edu/models). Another stated goal is ‘‘to use the CCSM to address important scientific questions about the climate system, including questions pertaining to global change and interdecadal and interannual variability’’. Moreover, the develop- Fig. 1 a Mean greenhouse effect (GH) in over 5 years of observations and b 10 years of CCSM. The colors and contours represent GH trapping in W/m2. The color scale is the same on each panel and the contours are drawn at 25 W/m2 intervals starting with 125 W/m2. Vectors represent mean 850 hPa wind direction and velocity (a reanalysis, b CCSM) with the reference maximum velocity magnitude shown over Asia. Here, as in other figures, winds vectors are interpolated on a 5 · 5 grid, for easier visualization 207 ers encourage the climate community to ‘‘diagnose and suggest possible avenues to improve the component models so that the full CCSM can better simulate coupled atmosphere–ocean variability on intraseasonal, seasonal and interannual time scales’’. Researchers should be able to use the CCSM for ‘‘predictability and sensitivity studies to further understanding of the nature and predictability of interannual variations of the climate system and their importance in the dynamics of longer-term climate variations and climate response to external forcing’’ (Blackmon et al. 2001). Plentiful data from a 300-year control simulation that has reproduced stable surface temperatures without artificial flux adjustments as well as a climate change run have been made freely available on the Internet. We therefore chose the CCSM for its quality, availability and openness to scrutiny by the wider climate research community. From the 300-year control simulation, we chose the last 10-year period, which showed pronounced interannual variability. As in the observations, we use modeled SST and OLR to compute GH, and analyze it together with modeled LHF and 850 hPa winds. 3 Means, variances and local correlations Mean observed and modeled patterns of GH are remarkably similar (Fig. 1). Tropical GH is strongest over the warm pool and the inter-tropical and south Pacific convergence zones (ITCZ and SPCZ, respectively). In general, the strongest GH follows convection in regions of wind (and moisture) convergence. Low values of GH and strong winds are found in the 208 Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale subtropical highs. The modeled GH is too low over the Indian Ocean and in the Atlantic ITCZ as well as in the equatorial central and eastern Pacific. The modeled SPCZ appears to extend too far into the tropical southcentral Pacific suggesting the existence of a double ITCZ in the model. Overall, however, the qualitative and quantitative agreement between CCSM and observations is within allowable limits. The major discrepancy between model and observations is that the mean model greenhouse appears to be of roughly equal magnitude in both hemispheres, while in the observations, Northern Hemisphere oceans are marked with a stronger mean GH. This observed result is consistent with precipitation observations (see e.g., Gershunov and Michaelsen 1996, their Fig. 4a). Mean observed and modeled LHF fields are displayed in Fig. 2. The top panel suggests that regions of strong GH are typically regions of weak evaporation and LHF while the largest sources of moisture are found in subsidence zones characterized by strong winds. Strong LHF is also observed in the western boundary currents, e.g., the northward branches of the Kuroshio and the Gulf Stream. The weakest LHF is in the eastern margins of the equatorial Pacific and Atlantic Oceans. Modeled mean LHF (Fig. 2b) only roughly resembles the observed. The largest regions of evaporative cooling and moisture sources are located away from the equator Fig. 2a, b Same as Fig. 1, but for latent heat flux (LHF). Contours start at 25 W/m2 and the region of lowest LHF is the equatorial Pacific cold tongue, which appears to be too long in the model. The Gulf Stream and the Kuroshio Currents are marked with remarkably high LHF, higher than observed. Modeled winds are also stronger then observed in the northwestern Pacific. The model is characterized by too much evaporation in general, especially in the equatorial western Pacific and Indian oceans as compared with observations. Moreover, extensive tropical Pacific moisture sources tend to be located in the central and western tropical Pacific and, therefore, are not as clearly spatially de-coupled from regions of strong GH in CCSM as they are in the observations. From the broadest point of view, observations show that, on average, the Southern Hemisphere produces more moisture as evaporation. In the model, however, both hemispheres are marked with comparably high LHF. Despite being small, the observed interhemispheric gradient is a robust feature of various available LHF observational data sets (Kubota et al. 2003), especially in the Pacific. The model, however, displays the opposite interhemispheric LHF gradient in the Pacific, e.g., less evaporation in the Southern than in the Northern Hemisphere. Figures 3 and 4 quantify the standard deviation in observed and modeled GH and LHF total fields. The largest GH variability occurs in the Asian – Australian Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale monsoon regions and, by extension, around the western Pacific warm pool, the most spatially extensive region of strong GH. Other regions of strong variability are found around central America, equatorial Atlantic, the north and south Atlantic convergence zones, and the Guinea Basin. A distinct region of high observed GH variability is located in the central equatorial Pacific east of the date line (Fig. 3a). Modeled GH variability is less representative of reality than the modeled mean GH. Modeled variability is too strong in the tropical seasonal convergence zones. This high variability extends all the way around the tropics on both sides of the Equator, even in the central and eastern Pacific, further suggesting the presence of a double seasonal ITCZ in the model. Equatorial GH variability is lower, as in the observations, but still too high everywhere except the central– eastern tropical Pacific where it is too low. Curiously, while mean GH is reproduced better by CCSM than the GH variability, LHF variability is better reproduced than the mean field. Both in the observations and the model, the strongest LHF variability is found in the northward branches of the Kuroshio and the Gulf Stream Currents (Fig. 4). High LHF variance is also apparent in the tropical North Atlantic in a band stretching southwestward from West Africa to Brazil. There is a gradient of increasing variability away from Fig. 3 a Standard deviation of GH in observations and b CCSM. The color scale is the same on each panel. Contours are drawn at 5 W/m2 intervals. Vectors represent the standard deviation of 850 hPa wind u and v components. Reference maximum standard deviation magnitude is shown over Asia 209 the Equator and also westward across ocean basins. The Northern Hemisphere oceans are marked with appreciably stronger LHF variability then the Southern Hemisphere oceans. All these features of LHF variability appear in the observations and model alike. The model differs from observations mainly in having too much LHF variability in the equatorial western Pacific and Indian Oceans and around Japan; and too little in the southeast tropical Atlantic and Pacific, the southwestern Indian Ocean and the central equatorial Pacific. To see that GH and LHF are inversely related in time, consider Fig. 5a, which shows the observed local temporal correlation between monthly means of GH and LHF over the 5-year observational period. Correlations are negative everywhere except for the north- and south-eastern Pacific, regions of low average GH as well as low variability in both GH and LHF fields; and in the central equatorial Indian Ocean, around Indonesia, the far eastern Pacific and Atlantic Oceans, regions of low mean and variance in the observed LHF field. These areas are geographically small and the largest positive correlations there are 0.3. Most of the tropics show negative correlations reaching below –0.75 in the southwestern Indian Ocean, northwestern Pacific, central Pacific and western Atlantic, just south of the equator. However, strong negative correlations are 210 Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale Fig. 4 Same as Fig. 3, but for LHF found practically everywhere the standard deviation of both GH and LHF is high and where the mean values are considerable as well. In most other areas, correlations are lower, but predominantly negative. Modeled GH–LHF correlations (Fig. 5b) are negative in regions where either GH or LHF variability is very strong. However, these regions are less geographically extensive, while regions of positive correlations are more extensive and feature stronger correlations than in the observations. This propensity to more positive correlation is clearly visible in the contrast between the observed and modeled local correlation histograms or empirical probability density functions (PDFs: Fig. 5c). On first approximation however, the general modeled patterns are not too un representative, except in the central equatorial Pacific where modeled correlations tend to be positive, while the observed ones are strongly negative. The general picture of GH and LHF means and variances gives a sense of spatial de-coupling between LHF and GH, especially in the observations. In the CCSM, the mean GH is well reproduced, the variability is not; LHF presents the opposite story. However, to first approximations, the comparison between modeled and observed means and variances largely validates the model. Local anti-correlation points to strong but nonlocal coupling between the sources of moisture and its atmospheric sinks, as discussed by G98. The observational results validate the results of G98, obtained using different data sources on a shorter and different period of record (July 1987–February 1990). Model-derived local correlations tend to be less negative, but generally display correct large-scale patterns, except in the equatorial central and eastern Pacific region. Based on these routine diagnostics, nothing further can be said about the ‘‘coupled’’ nature of GH and LHF. Clearly, a thoroughly non-local analysis is needed to describe this coupled variability. 4 Coupled GH–LHF variability 4.1 Analysis procedure Canonical correlation analysis (CCA: Barnett and Preisendorfer 1987; Bretherton et al. 1992) is a multivariate statistical technique used to summarize patterns of coupled variability in two fields of variables. Here, locations in the spatial fields (grid cells) are the relevant variables. The ordering of variables is not weighted, so, for geographically distributed fields, no preference is given to local coupling above temporal relationships over great distances. Since there are many more Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale 211 Fig. 5 Local temporal correlations between GH and LHF in a the observations and b CCSM. Positive contours are solid, negative are dashed, the zero contour is heavily thickened and the ±0.5 and ±0.75 contours are lightly thickened. The thin contour interval is 0.1. c Shows the contrast between observed and modeled local correlations (see text for details) variables in space than observations in time (3496 observed ocean grid cells and 60 months of observations, 2830 grid cells and 120 months of CCSM data) and since each geographical field is highly spatially autocorrelated, we reduce the dimensionality of each data set by prefiltering it separately with principal component analysis (PCA). Ten leading PCs are retained to summarize the variability in each field. Ten leading PCs account for 82.5% and 72.4% of the total variability in observed GH and LHF, respectively; 83.1% and 74.1% for modeled GH and LHF. CCA then looks for pairs of GH and LHF patterns (i.e., linear combinations of these sets of 10 GH PCs and 10 LHF PCs) whose temporal evolution is maximally correlated. After finding the leading canonical correlation pair, CCA searches for the next pair orthogonal to the first, and so on. For reasons that will become obvious, we only consider the leading canonical correlation pair (CC1: the leading coupled mode). Exactly the same analysis is performed for observations and for model data. 212 Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale 4.2 Seasonal cycle The leading coupled GH–LHF mode in observations is displayed in Fig. 6. Its time evolution is the seasonal cycle with a maximum in January and a minimum in July. The spatial patterns of the coupled seasonal mode in GH (Fig. 6b) and LHF (Fig. 6c) are presented in units of W/m2, with the same color scale and contouring convention as the standard deviation in Figs. 3a and 4a, respectively. Phase information is revealed via solid (in Fig. 6 The leading coupled mode (CC1) of observed GH and LHF. Time series of CC1 evolution for the GH (solid) and LHF (dashed) components of the coupled leading mode. Vertical long- (short-) dashed lines are drawn to mark January (July) positions. GH and LHF time series are then correlated with their respective raw fields to obtain a meaningful representation of the CC1 spatial patterns. The spatial patterns of CC1 b GH and c LHF are displayed. The squared correlations representing the proportion of local variance explained by CC1 are multiplied by the local standard deviation and displayed on the same color scale as the standard deviations in Figs. 3a and 4a for GH and LHF, respectively. The contours represent the same information but are drawn solid where the original correlations are positive, and dashed where they are negative. Phase information is thus represented by contour line type. As in Figs. 3 and 4, contours are drawn at 5 W/m2 intervals. The zero contour is thickened. Arrows represent anomalous winds associated with the CC1 of GH and LHF represented as correlations between a CC1 temporal evolution and the u and v components of the 850 hPa winds. In the present situation, these vectors can be interpreted as January–July winds phase) and dashed (180 out of phase) contours. This spatial quantification of the seasonal cycle can be interpreted with respect to the total variability of the individual raw fields as explained next. The seasonal cycle (CC1) explains most of the variance in the centers of variability in raw GH (compare Figs. 6b and 3a) and LHF (compare Figs. 6c and 4a) fields everywhere except for the equatorial Pacific. The proportion of GH and LHF standard deviation explained by CC1 in highly variable areas away from the equatorial Pacific is at least Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale half and much more in the primary centers of variability outlined in Figures 3 and 4. Thus, for example, over 80% of the total GH standard deviation is due to the seasonal cycle in the Bay of Bengal, Timor and Arafura Seas (North of Australia), South China Sea, Sea of Japan and around Central America. In LHF, the seasonal cycle accounts for over 80% in the Kuroshio and Gulf Stream Current regions as well as the north and south tropical Atlantic Ocean centers of action. Somewhat less, but still impressive proportions of variability are explained by CC1 in most other regions of high variability in GH and LHF (e.g., eastern Pacific 315–20N and S). The equatorial Pacific is an exception, this area is considered in detail in the following section. Phase information (contours) on Fig. 6b, c clearly shows that moisture sources in the subtropical highs are enhanced equatorward in the winter hemisphere when they provide moisture for the opposite summer–convective, enhanced-GH hemisphere. The seasonal cycles of GH and LHF are locally out of phase with each other. Contours on Fig. 6 show that GH is enhanced (diminished) in the summer (winter) hemisphere while the LHF seasonal cycle behaves in the opposite fashion. At first approximation, the intertropical belt is so partitioned into a convective (summer) hemisphere and an evaporative (winter) hemisphere. Anomalous winds tend to blow from the regions of high LHF (moisture sources) to the regions of high GH (atmospheric moisture sinks). The moisture evaporated from the Kuroshio Current region probably mostly fuels midlatitude cyclones. However, some of the moisture originating in the tropical northwestern Pacific may partially fuel convection in the SPCZ. Anomalous crossequatorial winds transport moisture from the tropical northwestern Pacific (Philippine Sea and further east) as well as from the seasonally weakened southeastern Pacific subtropical high to converge in the SPCZ in boreal winter. Conversely, in boreal summer, the enhanced GH in the northwestern Pacific convergence zone (NPCZ) is transported there from the northeastern Pacific high as well as by cross-equatorial flow from the winter hemisphere (to picture the boreal summer anomalous flow, mentally reverse the direction of the wind vectors). The northeastern Pacific moisture source (around Hawaii, see Fig. 2a) does not vary strongly with the seasons. The anomalous wind field over that region suggests that this moisture source is efficiently tapped by midlatitude cyclones in winter and by the NPCZ in summer. A mechanism involving large-scale atmospheric motion, the seasonally varying Hadley and Walker circulations, could explain much of the observed non-local coupling in GH and LHF. In the tropical-subtropical belt, evaporation in seasonal convergence zones is diminished by low-level convergence of moisture, which fuels convection and enhances the greenhouse effect. Away from areas of upward motion, GH is diminished and evaporation is enhanced by subsidence drying (Betts and Ridgway 1989), while evaporated moisture is removed via surface divergence. This moisture is carried 213 toward seasonal convergence zones by the seasonally varying trade winds. We next apply CCA to summarize the seasonal cycle in CCSM and find that the leading coupled GH–LHF mode in the coupled model also describes the seasonal cycle (Fig. 7a). This modeled seasonal cycle lags reality by one month with maxima typically in February and minima in August. As in the observations, regions of largest total GH and LHF variability (Figs. 3b and 4b) are largely explained by CC1. Zonally elongated regions of tropical off-equatorial Pacific and Atlantic GH variability are clearly seen to represent the seasonal migration of the modeled ITCZ into the summer hemisphere (Fig. 7b). The moisture sustaining this seasonal GH migration and warming originates in the evaporating and cooling waters of the winter hemisphere. However, the clean inter-hemispheric nature of CC1 in the observations is somewhat compromised in CCSM. The LHF seasonal cycle is not as inter-hemispheric in the model as it is in the observations. The meteorological equator is not as well defined in LHF CC1 (compare Figs. 6c and 7c). Moreover, only the centers of LHF variability are represented by the seasonal cycle with other extensive areas, especially in the Southern Hemisphere, exhibiting weakly opposite or no phase relationship with the dominant southern hemispheric seasonal cycle. In the observations, however, the seasonal cycle, to first approximation, is an inter-hemispheric seesaw. Most of the CCSM LHF variability is not accounted for by a single harmonic seasonal cycle. CCSM winds, however, show a consistent anomalous cross-equatorial transport from the winter to the summer hemisphere. In the Pacific, this transport is strongest and most zonally consistent. This is certainly related to the highly seasonally variable zonally elongated moisture source in the tropical Pacific 310N that obviously feeds the elongated zone of strong seasonal GH variability (modeled summer ITCZ) in the south tropical Pacific Ocean. It is worth noting here that the Gulf Stream is not as seasonally variable in both model and observations as the Kuroshio Current, both regions of very high total variability. These two regions exhibit very high annual mean LHF in reality and in the model. CCSM LHF values show a global maximum in the Gulf Stream: mean values above 200 W/m2, i.e., considerably higher than what is observed. Interestingly, LHF in both these intensely evaporating and highly variable regions, the Gulf Stream and the Kuroshio Current, is much less seasonally variable in the CCSM as it is in reality. The second canonical mode (CC2) in both observations and model (not shown) represents a seasonal cycle in quadrature with CC1 and explains the rest of seasonal variability with high loadings in low-variability regions that were not accounted for by CC1. Some of the dissimilarities between model and observations, especially the discrepancies in the spatial signatures of the seasonal cycle, stem from the differences in modeled and observed mean fields as well as differences in the uncoupled variability structure of the 214 Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale Fig. 7a–c Same as Fig. 6, but for CCSM individual fields. These dissimilarities lead to fundamentally similar coupled variability relative to the respective modeled and observed mean and variance fields. Other differences, such as the one-month setback in the modeled seasonal cycle and the generally weaker seasonal cycle in the modeled LHF field, could stem from thermodynamic problems in the model and can potentially lead to fundamental problems in moisture cycling and dynamics resulting in discrepancies between the modeled and observed global energy budgets. Notwithstanding these differences, on the whole, the leading coupled mode in the model and in the observations is the seasonal cycle characterized by generally similar behavior: a largely convective, enhanced GH, summer hemisphere and a moisture-supplying winter hemisphere. On the whole, CCSM does a fine job reproducing this general picture. 5 Interannual variability Beyond the seasonal cycle, the observations and CCSM data can be used to examine space-time scales of coupled interannual variability in GH and LHF. To do this, we Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale follow a similar approach as the one described to diagnose the seasonal cycle. 5.1 Data processing Before investigating interannual variability, we first removed the seasonal cycle from observed and modeled data. The double harmonic (annual and semiannual) cycles were fitted via least squares regression locally and subtracted from the GH, LHF and 850 hPa wind fields at each grid cell. In observations, interannual analysis including CCA was performed exactly as in the previous section on thus de-seasoned data. In the model, it was necessary to process the data further in order to focus on modeled interannual variability. The four-month exactly periodic cycle had to be removed from all model fields. This was done by locally fitting and subtracting a harmonic cycle. Because strong high-frequency variability still remained, the model deseasoned fields were then smoothed using a method of running medians known as 4(3RSR)2H (Tukey 1977). Smoothing was performed at each grid cell twice by smoothing once, computing the residuals from the smooth results, smoothing these and adding the two smoothed series together. Interannual signals are readily apparent in this super-smoothed model data. Fig. 8a, b Same as Fig. 3, but for standard deviation of the interannual GH and wind data 215 5.2 Variability and local correlations Before presenting the CCA results, we first briefly describe interannual variability of GH and LHF individually as well as their local temporal correlations in the de-seasoned observations and super-smoothed model data. Figures 8 and 9 present the interannual standard deviation of GH and LHF, respectively, for observations and CCSM. Figure 10 presents the local correlations. The observed GH is most variable in the areas of ENSO-related convection variability, i.e., the central and far western equatorial Pacific Ocean (Fig. 8a). This pattern is prominent in spite of the fact that ENSO activity was weak during 1992–1996, the observational period. An Indian Ocean center of interannual variability is also apparent. Modeled interannual variability largely resembles the observed, but only a single stronger center of action is found in the western equatorial Pacific. The Indian Ocean variability maximum is well reproduced in the model, but modeled variability is lower than observed in the central-eastern equatorial Pacific and in the equatorial Atlantic Oceans. Interestingly, the magnitude of interannual GH variability in the super-smoothed model data is roughly the same as that in the de-seasoned observations. We note again that the observational period was marked by particularly low 216 Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale ENSO activity, while the model period was chosen for its prominent interannual swings. The magnitude of ENSO SST variability, however, is known to be underestimated in CCSM (Blackmon et al. 2001). The level of modeled interannual LHF variability, in contrast, is considerably lower than the observed (Fig. 9). The magnitude of wind standard deviation is also much lower in the super-smoothed model data compared to observations. In the observations, aside from a variable region in the central equatorial Pacific, most of the variability is concentrated away from the Equator in loosely defined zonal bands between about 10 and 30N and S as well as in the Kuroshio and Gulfstream regions. The Hawaiian region and the southeastern Pacific Ocean are prominent centers of interannual LHF variability. The model also produces interannual variability in the northward branches of the Kuroshio and the Gulf Stream Currents as well as just to the south of the SPCZ, but elsewhere, model variability looks less realistic. A prominent horseshoe shaped region of enhanced variability extends from the western Pacific to the northeast and southeast. The observations of LHF could suffer from an overestimation of the variability in the high flux regions owing to the limitations of the satellite technique to account for cold air Fig. 9a, b Same as Fig. 4, but for standard deviation of the interannual LHF and wind data outbreaks and associated stratification effects on the fluxes (Bentamy et al. 2003). To which extent this might apply to the HOAPS climatology is unclear but we remind the reader that model – observations discrepancies are not only caused by the model and that possible limitations of the current data set should be borne in mind. Interannual wind variability in the observations certainly is stronger than that in the super-smoothed model data. Interestingly, while observed interannual wind variability is predominantly zonal at the equator (except in the Indian Ocean) and mostly meridional in the higher latitudes, the interannual modeled wind varies everywhere in both u and v directions on interannual time scales. Observed interannual GH–LHF local temporal correlations (Fig. 10a) show one coherent region of strong inverse relationship in the central equatorial Pacific stretching and diminishing towards the southeast. A large region around the Maritime continent is marked with blotches of strong negative correlation. Weakly positive correlations are observed in the eastern Pacific cold tongue and the tropical eastern Atlantic. Elsewhere in the tropics, correlations are predominantly negative but modest in the entire analysis domain with some Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale exceptions that lack spatial coherence. The modeled interannual GH–LHF local temporal correlations (Fig. 10b) also show predominantly negative values throughout the analysis domain, but their spatial distribution is noisier than in the observations. The most coherent spatial pattern of strong correlations is a swath of positive values along the equatorial Pacific band with well-defined maxima in the cold tongue and in the western equatorial Pacific Ocean. 5.3 Coupled variability Even during the observational period of weak ENSO activity, CCA efficiently isolates the ENSO signal in deseasoned GH–LHF co-variability. The temporal evolution of the coupled mode closely follows NINO3.4 (Fig. 11a). The spatial signature of CC1 is clearly the ENSO signal in GH (Fig. 11b). Enhanced GH in the central and eastern equatorial Pacific (lower in the far western Pacific) follows the typical El Niño signal in convection and precipitation (e.g., Wallace et al. 1998, Gershunov and Michaelsen 1996). Anomalous wind convergence into this area of enhanced GH can also be seen in Fig. 11b. Considered on a larger scale, the large Fig. 10a, b Same as Fig. 5, but for interannual GH and LHF data 217 coherent pattern of weakly positive and negative GH phases resembles Walker’s Southern Oscillation (Walker 1924). The LHF signature of this El Niño-related leading interannual coupled mode is a strong decrease in the central equatorial Pacific coincident with the strong local increase in GH there. Furthermore, a tail of decreased LHF stretches southeastward from this center of action during El Niño with weak decreases almost everywhere else in the central tropical Pacific. However, weak increases in LHF are observed in the equatorial Pacific cold tongue, where usually low LHF increases possibly due to warmer SST, as well as around the Maritime Continent and in the far western, off-equatorial Pacific tropical seasonal convergence zones, where usually low evaporation must increase during El Niño as a result of anomalous subsidence drying and low-level divergence. G98 obtained very similar ENSO signals in GH and LHF using different data covering part of one ENSO cycle (mostly the cold portion July 1987–1989). A similar mechanism can explain much of this locally inverse coupling observed on seasonal and interannual time scales. Low-level moisture convergence reduces anomalous LHF but enhances GH by fueling convection. This, in turn, promotes anomalous subsidence resulting in drying that reduces GH remotely and 218 Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale Fig. 11a–c Same as Fig. 6, but for observed interannual data. The green curve in a shows the observed NINO 3.4 SST evolution for reference together with low-level divergence enhances LHF. This remotely evaporated moisture is advected into the central equatorial Pacific Ocean during El Niño. Together, observed coupled patterns largely explain the equatorial structure of total interannual uncoupled variability in the GH and LHF (compare Figs. 11b, c with Figs. 8a and 9a, respectively). Of course, most of the central equatorial Pacific standard deviation in both GH and LHF is due to ENSO and is explained by CC1. Over 70% of the global maximum of GH standard deviation located in the central equatorial Pacific is explained by CC1. CC1 also explains most of the interannual GH variability in the entire Pacific region as well as considerable portions of LHF variability around the globe. Strong interannual LHF variability in the south-central Pacific, the southwestern Indian Ocean near Madagascar, and in the Gulf Stream remains largely unexplained by CC1. However, interannual LHF variability in many regions, even those far removed from the tropical Pacific, i.e., ENSO’s center of action, are related to CC1. Thus, substantial portions of the interannual LHF variability are explained by CC1 in the following highly variable regions: the southeastern Pacific (south of the cold tongue), the Kuroshio Current, Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale southeastern Indian Ocean (west of Australia), and parts of the tropical off-equatorial Atlantic Ocean. These apparent teleconnections are peripheral to the focus of this work, but they are interesting nonetheless. Admittedly, beyond the tropical Pacific, the significance of these relationships is questionable. The mechanisms of these LHF apparent teleconnections cannot be explored with 5 years of weak-ENSO activity data and are, in any case, beyond the scope of the current study. We suspect that, if these patterns are more than just sampling variability, they are mainly related to global circulation changes associated with ENSO. CC1 also explains most of the local correlation structure between GH and LHF (compare Figs. 11 and 10) in the equatorial Pacific. Throughout the ENSO cycle, GH and LHF are strongly inversely coupled in the central equatorial Pacific – the center of ENSO activity. Their coupling is weaker in the warm pool where GH decreases with decreased convergence and convection during El Niño and the LHF signal is split between an increase to the east and a decrease to the west. Positive in-phase coupling exists in the cold tongue where GH and LHF both increase with warmer SST during El Niño. Of course, it would help to consider an ENSOactive observational period to scrutinize and quantify the ENSO signal in GH–LHF co-variability, but the signal summarized here (Fig. 11) is certainly strong and coherent enough to make a comparison with the coupled model. The leading coupled mode of the interannual GH– LHF in the coupled model also evolves in line with CCSM’s ENSO (Fig. 12a). The ENSO pattern in modeled SST covers almost the entire equatorial Pacific stretching broadly from the south American west coast in a narrowing pattern to the western Pacific (see Blackmon et al. 2001). The CC1 signature in GH is weakly apparent in this entire region. However, the center of action in modeled GH CC1 is in its westernmost extremity, where a warm SST anomaly superimposed on climatologically warm SST enhances (decreases) modeled convection during warm (cold) phases of the modeled ENSO cycle. This ENSO pattern in GH appears to be out-of-phase with the central equatorial Indian Ocean GH. The GH signature of modeled CC1 is weaker than that in the observations, but still explains about 50% of total modeled interannual variability in its western equatorial Pacific center of action (compare Figs. 8b and 12b). The leading coupled mode signature in LHF is still weaker, but certainly coherent enough to describe a positive, in-phase relationship with GH in the western equatorial Pacific. Therefore, the strong positive local temporal correlations between modeled GH and LHF (Fig. 9b) result in large part from an incorrect local coupling of GH and LHF on ENSO time scales. When GH increases in the model with increased SST and convergence in the western equatorial Pacific, LHF also increases there, contrary to observed reality. A local source of moisture thus appears to exist in the model for the ENSO-enhanced 219 GH, whereas in observations, moisture sources are clearly remote. On longer than seasonal time scales, therefore, we have reason to believe that the CCSM does not correctly reproduce the observed GH–LHF dynamics. This means that we cannot use the current version of the CCSM for a detailed study of the interannual and longer time scale dynamics responsible for greenhouse warming and evaporative cooling of the ocean surface for natural or anthropogenically perturbed conditions. Several improvements planned for the CCSM may lead to a more realistic depiction of the GH–LHF interplay. These planned improvements include a more realistic parametrization of tropical convection; improved physical linkages between the parametrizations of convection, stratiform clouds, radiation and turbulence; addressing the ocean model’s sensitivities to representations of deep convection and boundary layer dynamics; and a more physically based specification of material fluxes through the sea surface (Blackmon et al. 2001). It is important to check whether the next version of the CCSM, as well as up-to-date versions of other available models, reproduce the observed GH–LHF coupling. 6 Summary and conclusions The aim here was to diagnose efficiently and quantify the coupled variability of the total (mainly water vapor and cloud) greenhouse effect and latent heat flux in observations and a coupled GCM. Low-level winds responsible for the transport of water vapor from surface sources to atmospheric sink areas were also considered. Due to the non-local nature of coupled relationships, diagnostic methods had to be applied in a non-local context. CCA was used to efficiently describe and summarize the salient modes of observed and modeled GH– LHF co-variability. We began with a general description and comparison of the observed and modeled GH and LHF mean fields, variability and local temporal correlation. Then, CCA results were presented in such a way as to quantify the proportion of total observed variability in the individual fields due to coherent and meaningful coupled modes. The nature of observational and model data allowed for the investigation of two modes: the seasonal cycle and ENSO. These modes were summarized in the context of coupled GH–LHF variability in observations and in a coupled model, the CCSM. Coupled modes of GH and LHF, derived to maximize correlation between temporal evolutions of patterns in the two fields, explain much of the total (uncoupled) variability in each field. This is somewhat less true in the CCSM than in the observations. Moreover, observations and model data clearly show that the coupling of GH and LHF is very strong on seasonal and interannual time scales and that seasonality and ENSO are responsible for much of the observed total variability in both GH and LHF (less so for the model). The time 220 Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale Fig. 12a–c Same as Fig. 6, but for CCSM interannual data. The green curve in a shows the modeled NINO 3.4 SST evolution for reference and space scales involved also clarify the mechanisms of the GH–LHF relationship and suggest the processes responsible for the interplay of these fields can, for the most part, be explained by large-scale atmospheric motion involving Hadley and Walker cell dynamics. There are important similarities between these mechanisms in observations and in the CCSM, especially on seasonal time scales, however there are also important differences between model and reality, especially due to processes operating on interannual (i.e., ENSO) time scales. Throughout the tropics and subtropics, remote moisture source regions sustain the greenhouse effect in regions of strongest GH via large-scale atmospheric dynamics. In fact, the spatial distributions of GH and LHF appear to be imbedded in the Hadley and Walker circulations and might reinforce these circulations by enhancing large-scale horizontal temperature and pressure gradients through local thermodynamics. The nonlocal nature of the GH–LHF coupling is manifested on at least seasonal and interannual time scales. For example, on seasonal time scales, when greenhouse warming intensifies in the summer hemisphere, evaporational cooling decreases there but increases in the opposite, i.e., winter, hemisphere. The model does a Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale reasonable job simulating this seasonal interplay even though modeled climatologies of GH and LHF are roughly symmetric around the equator; in the observations, on average, the northern tropics are more convective and are marked by higher GH, while the Southern Hemisphere evaporates more moisture per unit area. This mean interhemispheric gradient is especially clearly observed in the Pacific sector, where the model produces the opposite gradient in LHF. On interannual time scales, during El Niño, when strong convection and GH usually located over the western Pacific warm pool move eastward to the central Pacific, observed LHF decreases locally and in the subtropical highs, but increases almost everywhere else. Although the five observed years (1992–1996) were characterized by especially weak ENSO activity (Goddard and Graham 1997), our coupled analysis procedure was sensitive enough to pick out the ENSO pattern as the main mode of non-seasonal coupled variability. In contrast to the observations, the model provides a local source of moisture for ENSO-enhanced GH. A more interannually active observational time period is needed for a more detailed study of ENSO-related GH-LHF dynamics. Then again, a CGCM that correctly reproduces these dynamics can be used in a detailed study of interannual and longer time scale GH-LHF dynamics in the contexts of natural and anthropogenically perturbed climate variability. As far as the model ENSO goes, it is possible that incorrect LHF dynamics could be part of the reason why CCSM does not correctly simulate ENSO variability. In particular, the magnitude of ENSO SST variability in the CCSM is much too small (Blackmon et al. 2001). In the central and eastern equatorial Pacific, ENSO SST variability can be accounted for by heat advection and evaporative cooling (Niiler et al. submitted 2003). Evaporative cooling of the surface incorrectly positioned in the anomalous central Pacific warm pool during El Niño should dampen ENSO-related SST anomalies. The effect of this mechanism deserves further investigation, but again, a more suitable (i.e., ENSO-active) observational period is required for comparison with the model. A caveat of the present comparison concerns the observational data sets which are currently used. Indeed, the LHF data appears to suffer from underestimation of the mean in the tropics with respect to other recent satellite products (Kubota et al. 2003). While the main conclusions of this study stem from strong, clear signals, signals that were also reproduced in independent GH and LHF data covering a different time period (G98), one still needs to exercise caution when looking at the details of the discrepancies between the model integration and the present observational data set. In the near future, the expected better characterization of the observations’ limitations and the increase in their quality and degree of confidence will increase the usefulness of the present diagnostic and its ability to unravel models’ discrepancies in fine detail. Other available climatologies can be used to complement the HOAPS data set, e.g., the Goddard 221 Satellite-based Surface Turbulent Flux (Chou et al. 1997) and the Japanese Ocean Flux Data Sets using remote sensing observations (Kubota et al. 2002). These products could help bracket the uncertainty in the satellite observations of the LHF. Similarly, the use of recent CERES OLR observations could provide an alternative computation of the total sky GH allowing a stronger, more detailed characterization of models’ errors. After making these points, the clear and strong discrepancies between model and observations do not necessarily stem from an incorrect simulation of the individual fields of GH or LHF. Important discrepancies exist in the model relative to model climatology and variability. Such a fundamental mismatch between model and reality can lead to serious errors/biases in the modeled energy budget variance due to natural as well as to anthropogenic climate variability. Investigating these two highly integrated variables reveals the model’s limitations and points to a need for further detailed investigations of specific issues in the model. For instance, an analysis of the interannual variability in the wind-LHF relationship in both model and observations is one perspective for future work. Further study could partially explain the presently revealed discrepancies by isolating the respective contribution of the surface wind to the LHF variability with respect to the role of SST, transfer coefficients and surface moisture deficit. In the current analysis and interpretation of the CCSM results, it must be borne in mind that deviations from reality are model-specific. Similar investigations can and should be carried out with other available CGCMs as part of the coupled model intercomparison project. We hope to embark on such a comparison in the near future. Acknowledgements This work was realized during a visit of the first author (A.G.) to the Laboratoire de Météorologie Dynamique, Palaiseau, France under the auspices of visiting grants from both the Centre National de la Recherche Scientifique an the Institut Pierre Simon Laplace, Paris, France. Funding from National Science Foundation Grant ATM 99-01110 partially supported this work. The authors thank Drs. R. Kandel and M. Desbois for constructive discussions as well as J.-L. Monge and B. Bonnet for technical support. Two anonymous reviewers provided excellent comments and suggestions which improved the original text. References Barnett TP, Preisendorfer R (1987) Origins and levels of monthly and seasonal forecast skill for United States surface air temperatures determined by canonical correlation analysis. Mon Weather Rev 115: 1825–1850 Bentamy A, Katsaros KB, Mestas-Nuñez AM, Drennan WM, Forde EB, Roquet H (2003) Satellite Estimates of wind speed and latent heat flux over the global oceans. J Clim 16: 637–656 Betts K, Ridgway W (1989) Climatic equilibrium of the atmospheric convective boundary layer over a tropical ocean. J Atmos Sci 46: 2621–2641 Blackmon M and coauthors (2001) The Community Climate System Model. Bull Am Meteorol Soc 82: 2357–2376 Bretherton CS, Smith C, Wallace JM (1992) An intercomparison of methods for finding coupled patterns in climate data. J Clim 5: 541–560 222 Gershunov and Roca: Coupling of latent heat flux and the greenhouse effect by large-scale Chou S-H, Shie C-L, Atlas RM, Ardizzone J (1997) Air-sea fluxes retrieved from special sensor microwave imager data. J Geophys Res 102: 12,705–12,726 Cornejo-Garrido AG, Stone PH (1977) On the heat balance of the Walker circulation. J Atmos Sci 34: 1155–1162 Duvel J-P, Viollier M, Raberanto P, Kandel R, Haeffelin M, Pakhomov LA, Golovko VA, Mueller J, Stuhlmann R, Mueller J, and the International ScaRaB Scientific Working Group (2001) The ScaRaB–Resurs Earth Radiation Budget Dataset and first results. Bull Am Meteorol Soc 82: 1397–1408 Gershunov A, Michaelsen J (1996) Climatic-scale space-time variability of tropical precipitation. J Geophys Res 101: 26,297– 26,307 Gershunov A, Michaelsen J, Gautier C (1998) Large-scale coupling between the tropical greenhouse effect and latent heat flux via atmospheric dynamics. J Geophys Res 103: 6017–6031 Goddard L, Graham NE (1997) El Niño in the 1990s. J Geophys Res 102: 10,423–10,436 Grassl H, Jost V, Schulz J, Ramesh Kumar MR, Bauer P, Schluessel P (2000) The Hamburg Ocean–Atmosphere Parameters and Fluxes from Satellite Data (HOAPS): a climatological atlas of satellite-derived air-sea interaction parameters over the world oceans. 132 pp, Available from http://www.hoaps.zmaw.de/ Held IM, Soden BJ (2000) Water vapor feedback and global warming. Ann Rev Energy Env 25: 441–475 Inamdar AK, Ramanathan V (1998) Tropical and global scale interactions among water vapor, atmospheric greenhouse effect, and surface temperature. J Geophys Res 103: 32,177–32,194 Kalnay E, Kanamitsu M, Kistler R, Collins W, Deaven D, Gandin L, Iredell M, Saha S, White G, Woollen J, Zhu Y, Leetmaa A, Reynolds B, Chelliah M, Ebisuzaki W, Higgins W, Janowiak J, Mo KC, Ropelewski C, Wang J, Jenne R, Joseph D (1996) The NCEP/NCAR 40-Year Reanalysis Project. Bull Am Meteorol Soc 77: 437–472 Kubota M, Iwwasaka N, Kizu S, Konda M, Kutsuwada K (2002) Japanese ocean flux data sets with use of remore sensing observations (J-OFURO). J Oceanogr 58: 213–225 Kubota M, Kano A, Muramatsu H, Tomita H (2003) Intercomparison of various surface latent heat flux fields. J Clim 16: 670– 678 Lindzen R (1990) Some coolness concerning global warming. Bull Am Meteorol Soc 71: 288–299 Manabe S, Wetherald RT (1967) Thermal equilibrium of the atmosphere with a given distribution of relative humidity. J Atmos Sci 24: 241–259 Pierrehumbert RT (1995) Thermostat, radiator fins and the local greenhouse effect. J Atmos Sci 152: 1784–1806 Ramanathan V (1981) The role of ocean–atmosphere interactions in the CO2 climate problem. J Atmos Sci 38: 918–930 Reynolds R (1988) A real-time global sea surface temperature analysis. J Clim 1: 75–86 Schluessel P, Schanz L, Englisch G (1995) Retrieval of latent heat flux and longwave irradiance at the sea surface from SSM/I and AVHRR measurements. Adv Space Res 16: 10,107–10,116 Schulz J, Schluessel P, Grassl H (1993) Water vapour in the atmospheric boundary layer over oceans from SSM/I measurements. Int J Remote Sens 14: 2773–2789 Schulz J, Meywerk J, Ewald S, Schluessel P (1997) Evaluation of satellite derived latent heat fluxes. J Clim 10: 2782–2795 Smith E and coauthors (1996) Satellite-derived sea surface temperature data available from the NOAA/NASA Pathfinder Program, http://www.agu.org/eos_elec/95274e.html, American Geophysical Union Tukey JW (1977) Exploratory data analysis. Addison-Wesley, Reading, Massachusetts, USA Walker GT (1924) Correlation in seasonal variations of weather, IV, a further study of world weather. Mem Indian Meteorol Dept 24: 275–332 Wallace JM, Rasmusson EM, Mitchell TP, Kousky VE, Sarachik ES, VonStorch H (1998) The structure and evolution of ENSOrelated climate variability in the tropical Pacific – lessons from TOGA. J Geophys Res 103: 14,241–14,259 Zhu Y, Newell RE, Read WG (2000) Factors controlling uppertropospheric water vapor. J Clim 13: 836–848