16
409 MARCH 2004 AMERICAN METEOROLOGICAL SOCIETY | O cean surface fluxes of heat, moisture, and mo- mentum observed during field experiments show strong variability on temporal scales that range from the diurnal cycle to the life cycle of storms, and on spatial scales as small as that of an individual convective cloud. High-frequency variability (e.g., di- urnal, storm scale) in tropical air–sea fluxes has been hypothesized to influence intraseasonal and interannual variability of the monsoon (e.g., Webster et al. 1998) and the Pacific Ocean warm pool and El Niño (e.g., Sui and Lau 1997; Fasullo and Webster 2000). At high latitudes, large variations in surface fluxes and sea surface temperature are seen in re- sponse to storms, which impact the temperature, den- sity, and mixing in the upper ocean, further influenc- ing the atmospheric dynamics and thermodynamics. Storm-scale events have been hypothesized (e.g., Marshall et al. 1998; Nardelli and Salusti 2000) to be associated with ocean convection in the high-latitude water mass formation regions, contributing to deep water formation and the global ocean thermohaline circulation. Ocean mixing induced by tropical cy- clones might play an important role in driving the global ocean thermohaline circulation and, thereby, SEAFLUX BY J. A. CURRY, A. BENTAMY, M. A. BOURASSA, D. BOURRAS, E. F. BRADLEY, M. BRUNKE, S. CASTRO, S. H. CHOU, C. A. CLAYSON, W. J. EMERY, L. EYMARD, C. W. FAIRALL, M. KUBOTA, B. LIN, W. PERRIE, R. A. REEDER, I. A. RENFREW, W. B. ROSSOW, J. SCHULZ, S. R. SMITH, P. J. WEBSTER, G. A. WICK, AND X. ZENG Satellite-based datasets of surface turbulent fluxes over the global oceans are being evaluated and improved. AFFILIATIONS: CURRY AND WEBSTER—Department of Earth and Atmospheric Sciences, Georgia Institute of Technology, Atlanta, Georgia; BENTAMY—Institut Francais pour la Recherche et l’Exploitation de la Mer, Brest, France; BOURASSA AND SMITH—Center for Ocean–Atmospheric Prediction Studies, The Florida State University, Tallahassee, Florida; BOURRAS—Jet Propulsion Laboratory, California Institute of Technology, Pasadena, California; BRADLEYCSIRO Land and Water, Canberra, Australia; BRUNKE AND ZENGDepartment of Atmospheric Sciences, University of Arizona, Tucson, Arizona; CASTRO, EMERY, AND REEDER—Department of Aerospace Engineering Sciences, University of Colorado, Boulder, Colorado; CHOU—NASA GSFC, Greenbelt, Maryland; CLAYSON AND EYMARDCETP/IPSL/CNRS, Velizy, France; FAIRALL AND WICK—NOAA ETL, Boulder, Colorado; KUBOTA—School of Marine Science and Technology, Tokai University, Orido, Shimizu, Shizuoka, Japan; LINNASA Langley Research Center, Hampton, Virginia; PERRIEBedford Institute of Oceanography, Dartmouth, Nova Scotia, Canada; RENFREW—Physical Sciences Division, British Antarctic Survey, Cambridge, United Kingdom; ROSSOW—NASA GISS, New York, New York; SCHULZ—Meteorological Institute, University of Bonn, Bonn, Germany CORRESPONDING AUTHOR: Judith A. Curry, Dept. of Earth and Atmospheric Sciences, Georgia Institute of Technology, Atlanta, GA 30332 E-mail: [email protected] DOI: 10.1175/BAMS-85-3-409 In final form 14 October 2003 ©2004 American Meteorological Society

SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

409MARCH 2004AMERICAN METEOROLOGICAL SOCIETY |

O cean surface fluxes of heat, moisture, and mo-mentum observed during field experimentsshow strong variability on temporal scales that

range from the diurnal cycle to the life cycle of storms,and on spatial scales as small as that of an individualconvective cloud. High-frequency variability (e.g., di-urnal, storm scale) in tropical air–sea fluxes has beenhypothesized to influence intraseasonal andinterannual variability of the monsoon (e.g., Websteret al. 1998) and the Pacific Ocean warm pool andEl Niño (e.g., Sui and Lau 1997; Fasullo and Webster2000). At high latitudes, large variations in surface

fluxes and sea surface temperature are seen in re-sponse to storms, which impact the temperature, den-sity, and mixing in the upper ocean, further influenc-ing the atmospheric dynamics and thermodynamics.Storm-scale events have been hypothesized (e.g.,Marshall et al. 1998; Nardelli and Salusti 2000) to beassociated with ocean convection in the high-latitudewater mass formation regions, contributing to deepwater formation and the global ocean thermohalinecirculation. Ocean mixing induced by tropical cy-clones might play an important role in driving theglobal ocean thermohaline circulation and, thereby,

SEAFLUXBY J. A. CURRY, A. BENTAMY, M. A. BOURASSA, D. BOURRAS, E. F. BRADLEY, M. BRUNKE, S. CASTRO,

S. H. CHOU, C. A. CLAYSON, W. J. EMERY, L. EYMARD, C. W. FAIRALL, M. KUBOTA, B. LIN,W. PERRIE, R. A. REEDER, I. A. RENFREW, W. B. ROSSOW, J. SCHULZ, S. R. SMITH,

P. J. WEBSTER, G. A. WICK, AND X. ZENG

Satellite-based datasets of surface turbulent fluxes over the global oceans

are being evaluated and improved.

AFFILIATIONS: CURRY AND WEBSTER—Department of Earth andAtmospheric Sciences, Georgia Institute of Technology, Atlanta,Georgia; BENTAMY—Institut Francais pour la Recherche etl’Exploitation de la Mer, Brest, France; BOURASSA AND SMITH—Centerfor Ocean–Atmospheric Prediction Studies, The Florida StateUniversity, Tallahassee, Florida; BOURRAS—Jet Propulsion Laboratory,California Institute of Technology, Pasadena, California; BRADLEY—CSIRO Land and Water, Canberra, Australia; BRUNKE AND ZENG—Department of Atmospheric Sciences, University of Arizona, Tucson,Arizona; CASTRO, EMERY, AND REEDER—Department of AerospaceEngineering Sciences, University of Colorado, Boulder, Colorado;CHOU—NASA GSFC, Greenbelt, Maryland; CLAYSON AND EYMARD—CETP/IPSL/CNRS, Velizy, France; FAIRALL AND WICK—NOAA ETL,Boulder, Colorado; KUBOTA—School of Marine Science andTechnology, Tokai University, Orido, Shimizu, Shizuoka, Japan; LIN—

NASA Langley Research Center, Hampton, Virginia; PERRIE—Bedford Institute of Oceanography, Dartmouth, Nova Scotia,Canada; RENFREW—Physical Sciences Division, British AntarcticSurvey, Cambridge, United Kingdom; ROSSOW—NASA GISS,New York, New York; SCHULZ—Meteorological Institute,University of Bonn, Bonn, Germany

CORRESPONDING AUTHOR: Judith A. Curry, Dept. of Earthand Atmospheric Sciences, Georgia Institute of Technology,Atlanta, GA 30332E-mail: [email protected]: 10.1175/BAMS-85-3-409

In final form 14 October 2003©2004 American Meteorological Society

Page 2: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

410 MARCH 2004|

in regulating climate (Emanuel 2001). Hence, evenlonger time-scale climate issues may be influenced byhigh-frequency interactions between the ocean andthe atmosphere.

A major issue in ocean modeling is the resolutionrequired in the spatial–temporal surface forcing andthe level of detail required for the parameterizationof the interfacial processes. While it is clear that di-urnal cycle forcing must be included to simulate ac-curately the upper ocean, it is not clear what horizon-tal scales of forcing are needed to simulate correctlythe major aspects of the ocean circulation that feedback onto the atmosphere. Different horizontal scalesof forcing may be important in different regimes,depending on the dominance of surface forcing byeither momentum or surface buoyancy fluxes. In gen-eral, using monthly mean winds and fluxes providessimulations that are too cold in the eastern Pacific andtoo warm in the western Pacific (e.g., Hayes et al.1989). Apparently, a shorter time-averaging periodfor the surface fluxes and/or a proper averaging overnonlinear behavior is needed to simulate the correctocean climatology. Accurate observations of high-resolution surface flux components are also neededto identify and resolve problems, such as the modeldrift that characterizes most coupled models (e.g.,Josse et al. 1999).

The need for high-resolution, accurate surfacefluxes of heat, water vapor, and momentum over theglobal ocean has been articulated by numerous groupswithin the global climate community, including theWorld Climate Research Programme (WCRP)Working Group on Air–Sea Fluxes (WGASF), theWCRP Global Energy and Water Cycle Experiment(GEWEX) Radiation Panel, and the Climate Varia-tions (CLIVAR) Science Steering Group. TheGEWEX Radiation Panel and the U.S. CLIVAR Com-mittee have established a goal of 1o spatial resolution,3–6-h time resolution, and accuracy of 5 W m-2 forindividual components of the surface heat budget.

Surface analyses from numerical weather predic-tion (NWP) models can provide one source of high-resolution surface fluxes. However, there are a num-ber of established problems with NWP fluxes, theprimary one being that these fluxes are not the prod-uct of the model analysis cycle, but are produced bythe forecast cycle that is more dependent on the physi-cal parameterizations used in the particular modeland not strongly constrained by observations. Atpresent, surface turbulent flux analyses from NWPmodels do not show sufficient variability and can havesubstantial biases in certain regions, notably, the high-latitude oceans (Renfrew et al. 2002), the Gulf Stream

(Moore and Renfrew 2002), and the Tropics (Smithet al. 2001). In NWP models, sea surface temperature(SST) is specified from satellite observations with aneffective resolution of >100 km and is averaged overa 5-day period. Most models, including that of the Eu-ropean Centre for Medium-Range Weather Forecasts(ECMWF), simulate lower-than-observed kinetic en-ergy at spatial scales less that 1000 km. The tropicalMadden–Julian oscillation is also poorly simulated. Itis hypothesized here that one of the major reasons forthe energy deficit and the poor simulation ofintraseasonal variability is that fluxes occurring athigh frequency and small scales (along with corre-sponding variations in skin sea surface temperature)are not correctly accounted for. These model deficien-cies appear to be related to the interfacialparameterizations (e.g., surface turbulent fluxes andskin SST), the cloudy boundary layer, and convectiveprocesses. Moreover, the specified SST cannot re-spond to changing atmospheric conditions, alteringthe interaction of the atmosphere and ocean surface.Because of these deficiencies, NWP surface fluxes donot presently provide the needed accuracy and reso-lution for reliable air–sea flux fields.

Satellite-based observations of air–sea fluxes pro-vide an alternative to NWP fluxes for global fluxfields. The surface momentum flux and precipitationcan be related relatively directly to satellite observa-tions. The surface radiation fluxes and sensible andlatent heat fluxes are determined using a radiationtransfer model and bulk turbulent flux model withsatellite-derived input variables. The emphasis to dateof satellite-derived flux datasets has been on monthlymean values, such as the GEWEX Global Precipita-tion Climatology Project (GPCP) and Surface Radia-tion Budget (SRB) Program, although efforts areunderway to produce higher-resolution datasets.Detailed evaluation and intercomparison studieshave been conducted for the SRB (Gupta et al. 1999)and GPCP (Ebert et al. 1996).

Several efforts are underway to prepare ocean sur-face turbulent flux datasets from satellites using bulkturbulent flux models. Bulk flux models link turbu-lent fluxes to mean values of surface temperature,wind, and surface air temperature and humidity, eachof which is determined from the satellite. The Ham-burg Ocean Atmosphere Parameters and Fluxes fromSatellite Data (HOAPS) dataset has been prepared atgrid resolutions of 1° ¥ 1° and 2.5° ¥ 2.5° as pentadand monthly fields for July 1987–December 1998(Schulz et al.1997). The HOAPS group is presentlyworking to produce a higher-resolution product.Another global flux product, the Goddard Satellite-

Page 3: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

411MARCH 2004AMERICAN METEOROLOGICAL SOCIETY |

Based Surface Turbulent Fluxes version 2 (GSSTF-2),described by Chou et al. (2003), has a resolution of 1o

and 1 day for July 1987–December 2000. The Japa-nese Ocean Flux Data Set with Use of Remote Sens-ing Observations (J-OFURO) (Kubota et al. 2002) isa monthly surface flux product with 1o resolution.Bentamy et al. (2003) have produced a global fluxproduct with 1o and weekly resolution for the periodof 30 September 1996 through 29 June 1997. Higher-resolution regional flux products include the Tropi-cal Ocean and Global Atmosphere (TOGA) CoupledOcean–Atmosphere Response Experiment (COARE)region (Curry et al. 1999), the tropical Pacific Ocean(Jones et al. 1999), the global tropical oceans (Lin et al.2001; Schlussel and Albert 2001), and the Mediterra-nean Sea (Bourras 2002b).

To date, there has been no effort to systematicallycompare and evaluate the satellite-derived ocean sur-face turbulent flux products, although each of the fluxdatasets has been evaluated using a limited amountof in situ data. Recently, Kubota et al. (2003) com-pared the zonal and annual average surface latent heatfluxes from National Centers for Environmental Pre-diction (NCEP) and ECMWF analysis, three differ-ent satellite flux products (Kubota et al. 2002; Chouet al. 2001; Schulz et al. 1997), and the Comprehen-sive Ocean–Atmosphere Data Set (COADS) Projectflux climatology derived from ship and buoy obser-vations (DaSilva et al. 1994). The results of this com-parison are summarized in Fig. 1. The differencesamong the three satellite flux datasets are approxi-mately the same magnitude as the differences amongthe NCEP–National Center for Atmospheric Re-search (NCAR) reanalysis (NRA1), ECMWF 15-yrreanalysis (ERA15), and COADS fluxes. The latentheat fluxes of J-OFURO and GSSTF-1 are similar, butHOAPS is smaller than J-OFURO and GSSTF in thetropical regions; HOAPS is closest to the COADSdataset. The satellite datasets show greater discrepan-cies with the COADS and NWP datasets in the South-ern Hemisphere relative to the Northern Hemisphere,including low temporal correlations (Kubota et al.2003). This suggests that COADS and NWP may beaffected by the lack of observations in the SouthernHemisphere, although increased assimilation of sat-ellite data by NWP models is expected to reduce thebias in Southern Hemisphere NWP products.

Although the COADS fluxes are the generally ac-cepted climatology, it is likely that this product is nomore accurate than the NWP or satellite climatolo-gies, because of the space/time sampling problems in-herent in any global dataset based on ship and buoyobservations, and the reliance on the same bulk mod-

els to calculate the fluxes. Studies of the uncertaintiesof surface fluxes based on datasets like COADS indi-cate that sampling error is as important as measurementerror (Weare 1989). The need to adjust the COADSfluxes to force ocean models has been thoroughly dis-cussed by WGASF (2001). WGASF (2001) comparedall available fields of zonal mean fluxes and evaluatedthe implied heat and moisture transports. It is clearfrom these comparisons that there are large discrepan-cies in the different flux products, and that even climato-logical estimates of the fluxes must be questioned.

Both the NWP and satellite products are capable ofproviding approximately the same resolution in surfaceflux products, which far exceeds the sampling of con-ventional sources. Without the high-resolution satelliteflux product, NWP products cannot be adequatelyevaluated in terms of scales of variability of the fluxes.To improve our understanding and determination ofocean surface turbulent flux products, the GEWEX

FIG. 1. Comparison of zonally averaged surface latentheat fluxes for Jan: (a) COADS (black), NRA1 (blue),and ERA15 (orange); (b) COADS (black), J-OFURO(red), HOAPS (green), and GSSFT1 (purple). FollowingKubota et al. (2002).

Page 4: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

412 MARCH 2004|

Radiation Panel has initiated the SEAFLUX Project,to address the following specific issues:

• What is feasible in terms of the time–space reso-lution and length-of-time series for a global oceansurface turbulent flux dataset?

• Can we produce a high-resolution ocean surfaceturbulent flux dataset using satellites that are betterthan conventional climatology and NWP products?

• What are the best methods for developing a high-resolution ocean surface turbulent flux dataset?

• How do the different flux products perform in thetarget applications (e.g., constraining the budget andmean transports of heat and freshwater in the glo-bal ocean, diagnosing regional and time variationsof the coupled atmosphere–ocean system, evaluat-ing the surface fluxes in coupled atmosphere–oceanmodels and weather forecasting models, and provid-ing surface forcing for ocean models)?

An overview is given in this paper of the method-ology being used by SEAFLUX, including theSEAFLUX Intercomparison Project. Further infor-mation on SEAFLUX can be obtained online athttp://curry.eas.gatech.edu/SEAFLUX/index.html. Ad-ditional background on air–sea fluxes is given in thecomprehensive report written by WGASF (2001).

STRATEGY OVERVIEW. Because other interna-tional efforts are focused on surface radiation fluxes(SRB) and precipitation (GPCP), the focus ofSEAFLUX is on the surface turbulent fluxes (sensibleheat, evaporation/latent heat, and momentum), whilerecognizing that most applications require an inter-nally consistent flux dataset that also includes the ra-diative fluxes and precipitation. The general methodthat has been adopted to determine the surface tur-bulent fluxes from satellite observations is to use abulk surface turbulent flux model to calculate thefluxes, using satellite-derived values of sea surfacetemperature, surface winds, and surface air tempera-ture and humidity as input. Surface momentum fluxescan be determined directly from scatterometer data(Weissman and Graber 1999). While a long-term glo-bal flux dataset is desirable, particularly for studies ofinterannual variability, it was determined to be fea-sible only to go back as far as 1987, given the avail-ability of appropriate satellite sensors.

The SEAFLUX project has the following elements:

• an extensive library of in situ datasets from re-search ships and buoys for validation of the globalflux products,

• a library of satellite datasets collocated with the insitu datasets, covering an area of approximately200 km around each in situ point,

• a library of NWP surface flux analyses collocatedwith the in situ datasets,

• the evaluation and improvement of bulk turbulentflux models,

• the production of a high-resolution skin sea sur-face temperature product from satellite observa-tions that includes the diurnal cycle,

• the evaluation and improvement of methods todetermine surface air temperature and humidityfrom satellites,

• the production and evaluation of global high-reso-lution satellite-derived surface turbulent fluxes,and

• the evaluation of the global flux products in thecontext of applications (e.g., forcing ocean mod-els, partitioning of heat transport in the atmo-sphere and ocean) and assessment of the senstivityof these studies to random and bias errors in theflux products.

The SEAFLUX Intercomparison Project uses dataselected from the period since July 1987. Cases ear-lier than 1987 are not used because of the absence ofSpecial Sensor Microwave Imager (SSM/I) data andbecause of improvements to observational strategies.The dataset is assembled in a manner that emphasizesease of use. All data are in a common ASCII formatwith no packing. Metadata are included for each insitu dataset, including a brief description of the instru-ments, journal references that describe the data anderror characteristics, quality control, and data format-ting information. A variety of satellite datasets are in-cluded to encourage scientists to try using other sat-ellite data sources. The in situ and collocated satelliteand NWP datasets have been assembled, and can befound online at http://curry.eas.gatech.edu/SEAFLUX/intercomparison-cg.html.

Discrepancies between the different satellite fluxproducts and their errors may be associated with sam-pling, the bulk turbulent flux model, and the accu-racy of the input variables for the flux model (e.g.,winds, SST, surface air temperature, and humidity).To address these discrepancies and errors, the follow-ing intercomparison projects have been formulated:

• bulk surface turbulent flux models against directturbulent flux measurements,

• satellite-derived and in situ winds,• satellite-derived and in situ bulk/skin sea surface

temperature,

Page 5: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

413MARCH 2004AMERICAN METEOROLOGICAL SOCIETY |

• pixel resolution satellite-derived and in situ sur-face air temperature and humidity, and

• global fields of surface turbulent fluxes for 1999.

This intercomparison framework promotes evalua-tion of existing flux products, formulation of new fluxproducts (determined by using different combina-tions of sources for input variables), and newly devel-oped methods to determine input variables. Theevaluation will not be conducted blindly; that is, par-ticipants will have access to the in situ data beforesubmitting their analyses for the intercomparison. Wehave opted not to conduct a blind intercomparison,because having the in situ data available will speedprogress on evaluating and improving the products.However, we are witholding data from the 1999 in situdeployments so that these data are not included inhigh-resolution evaluation and development of sat-ellite methods. The 1999 data will be reserved forevaluation of the global flux products for 1999. In thisway, there are some independent evaluation data thatwere not used in development of the methods.Comparison and evaluation of the global flux prod-ucts will be undertaken as follows:

1) pixel-level evaluations will be conducted using thelarge number of datasets, collected during 1999,to provide an independent evaluation of the de-rived fluxes; and

2) scales of variability will be compared and evalu-ated for the global dataset using an EOF analysis.

A key aspect of the evaluation is consideration ofthe utility of the Advanced Very High Resolution Ra-diometer (AVHRR) and SSM/I data (available forover a decade) versus the utilization of potentiallymore useful datasets, such as scatterometers andlower-frequency microwave observations, which haveonly recently become available.

In situ reference datasets. A variety of different typesof datasets can be used to evaluate the satellite-basedinput variables and derived fluxes, as well as NWPfluxes, with measurements obtained from buoys, re-search ships, and voluntary observing ships (for acomprehensive summary, see WGASF 2001). For thisstudy, we use only measurements that have been ob-tained from research ships and buoys. Although theamount of research-quality flux data is relatively smallwhen compared with the much larger dataset fromvoluntary observing ships, we prefer to use accuratemeasurements with carefully characterized errors toevaluate the satellite and NWP products.

FLUX MEASUREMENTSA general discussion of flux measurement issues isgiven in Fairall et al. (1997) and McGillis et al.(2001). Covariance fluxes require cross correlationof vertical velocity fluctuations with those of thehorizontal wind components (stress), temperature(sensible heat), and moisture (latent heat). A sonicanemometer is most commonly used to obtainthe three components of the wind vector (u¢¢¢¢¢, v¢¢¢¢¢,w¢¢¢¢¢) and the sonic temperature (T¢¢¢¢¢). A high-speedinfrared hygrometer is used to obtain specifichumidity (q¢¢¢¢¢). Typical fast-humidity technologiesinclude ultraviolet absorption, infrared absorption,or microwave refractive index. Open-oceanmeasurements are usually obtained from ships orbuoys, so platform motions must be removed.Inertial dissipation flux estimates are computedfrom the variance spectral density of u¢¢¢¢¢, T¢¢¢¢¢, and q¢¢¢¢¢in the inertial subrange of locally isotropic turbu-lence, which is usually at frequencies sufficientlyabove the wave-induced platform motions so thatcorrections are not needed. Flow distortion by theplatform structure also requires correction. Forwell-sited instruments on ships, the distortioneffects are thought to be negligible for scalar fluxes,but could be on the order of 10% for wind stress.

Covariance flux estimates are subject torandom sampling errors associated with atmo-spheric variability and other random errors causedby imperfect motion corrections or sensor noiseand drift. Systematic errors are caused by incor-rect sensor calibration, imperfect motion correc-tion, and flow distortion. If sensors are notphysically collocated, it is necessary to account forthe loss of correlation caused by the physicalseparation; such corrections are typically a fewpercent and in scale with the ratio of separation toheight above the surface. Inertial dissipation fluxestimates do not require motion corrections, andvariance estimates (i.e., variance spectra) alsohave smaller sampling variability than covariances.At low wind speeds, inertial dissipation estimateshave about one-third of the sampling uncertaintyof covariance estimates. However, inertial dissipa-tion estimates are subject to another major errorsource—uncertainty in the dimensionless struc-ture function parameter. Because of wave effects,dimensionless inertial dissipation functionsobtained over land are suspect for use near thesea surface.

For 50-min-averaged values, root-mean-squareerrors in direct turbulent flux measurements fromshipboard platforms are estimated to be 3 W m-----2

± 20% for covariance sensible heat flux, 5 W m-----2

± 20% for covariance latent heat flux, 0.015 N m-----2

± 30% for covariance wind stress, and 15% forinterital dissipation wind stress.

Page 6: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

414 MARCH 2004|

To date, we have identified in situ datasets fromthe following main sources:

• research cruises from the air–sea interaction groupat the National Oceanic and Atmospheric Admin-istration (NOAA) Environmental TechnologyLaboratory (ETL) (Fairall et al. 1997; Edson et al.1998),

• mooring observations (Weller and Anderson1996),

• skin sea surface temperature measurements(Emery et al. 2001),

• mid- and high-latitude observations from Germanresearch vessels (Bruemmer 1993),

• meteorological and flux data from World OceanCirculation Experiment (WOCE) cruises (Smithet al. 1996, 1999; available online at www.coaps.fsu.edu/woce/docs/qchbook/qchbook.htm),

• research cruises from French vessels (Weill et al.2003), and

• research cruises in the Indian Ocean by the Aus-tralian R/V Franklin (Godfrey et al. 1999).

All datasets have observations of position, bulk seasurface temperature, surface winds, air temperature,and humidity. For some cases, additional observationsare available of skin sea surface temperature, directturbulent flux measurements, or wave information.We also include observations of radiation fluxes andprecipitation, where available, because these param-eters are used in some of the bulk turbulent flux for-mulations. Some cases also include quality control in-dices.

The total number of deployments1 in the datasetcurrently numbers 54, providing approximately 290months of measurements. A total of 18 deploymentsinclude direct turbulent flux measurements—1 in-cludes wave information, and 16 include skin SSTdeterminations. A total of 21 deployments are fromthe Tropics (20oN–20oS), 14 deployments are fromthe subtropics (20oN–35oS), 14 are from themidlatitudes (35oN–55oS), and 5 are from high lati-tudes (55oN–80oS). All of the different datasets havebeen rewritten in a common format, using the sameunits, for ease of interpretation and use. The variablesin the in situ dataset are presented as 1-h averages,because we are comparing point measurements withsatellite observations or NWP estimates on a scale of50–100 km.

Satellite and NWP datasets. Satellite datasets includedin this study are chosen based upon numerous previ-ous research studies and recommendations from thefirst SEAFLUX workshop. We include both direct ra-diance and brightness temperatures, as well as satel-lite-derived products. Under this project, no attemptis made to improve upon the determination of param-eters that are being extensively addressed by othergroups, such as winds from scatterometers, verticalprofiles of atmospheric temperature and humidity,surface radiation fluxes, and precipitation. The follow-ing satellite radiance and brightness temperatures areprovided for each of the observational cases, at fullresolution of the observations, for the period since1987, unless otherwise noted:

• AVHRR Global Area Coverage (GAC): 4 km, 2–4times per day;

• SSM/I: pixel size from 69 km ¥ 43 km at 19 GHzto 15 km ¥ 13 km at 85 GHz, coverage 2 times perday for one satellite (since 1991 at least two satel-lites in orbit simultaneously);

• Tropical Rainfall Measuring Mission (TRMM) Mi-crowave Imager (TMI): pixel size from 63 km ¥ 68km at 10 GHz to 7 km ¥ 4.5 km at 85 GHz, wherecoverage is determined by precessing satellite withrepetition every 30 days (since December 1997).

The following satellite products are also provided:

• the NOAA operational bulk SST product(Reynolds and Smith 1995);

• passive microwave products (precipitable water,surface wind speed, rainfall) (Wentz 1997; version5), at the resolutions of SSM/I and TMI;

• NOAA Television Infrared Observation Satellite(TIROS) Operational Vertical Sounder (TOVS)temperature and humidity profiles—operationalproduct resolution is daily at 2.5o (effective), and3I product resolution is 2–4 times daily at 1o;

• scatterometer winds [from the National Aeronau-tics and Space Administration (NASA) Scatter-ometer (NSCAT) and Quick Scatterometer(QSCAT)] during 1996–97 and 1999–present, at25-km resolution;

• International Satellite Cloud Climatology Project(ISCCP) DX cloud products, at 30-km and 3-hresolution (Rossow and Schiffer 1999); and

• ISCCP-derived surface radiation fluxes, at 30-kmand 3-h resolution (Zhang and Rossow 1995).

All of these datasets are freely available either fromdata archives, via the Web, or are being provided by

1 Deployment is defined here as a nearly continous time seriesat a specific location (e.g., buoy) or a research vessel cruise.

Page 7: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

415MARCH 2004AMERICAN METEOROLOGICAL SOCIETY |

individuals associated with the project. The data areprovided at full resolution of the dataset, correspond-ing to the period and location of each in situ case, cov-ering a region of approximately 200 km x 200 kmaround the point location.

In addition to collocated satellite data and prod-ucts, we also include collocated data from the NCEPand ECMWF NWP analyses to include surface tem-perature, surface air temperature, humidity, winds,and the flux components. The NCEP analyses areobtained from the NCEP–NCAR reanalysis project.Variables from the surface flux data subset includewind speed at 10 m, air temperature and specific hu-midity at 2 m, skin temperature, downward radiationfluxes, and sensible and latent heat fluxes. Values areprovided globally at 6-h intervals (0000, 0600, 1200,1800 UTC). The ECMWF data are generated by theEuropean Center for Medium-Range Weather Fore-casts data assimilation system (information availableonline at www.ecmwf.int/). We use the ECMWFWCRP level-III-A Global Atmospheric Data Archive.The Advanced Operational Analysis Data Set includessurface variables, and the Supplementary Fields DataSet contains radiation and turbulent heat fluxes. Thetemporal coverage is 4 times daily at 0000, 0600, 1200,and 1800 UTC.

EVALUATION OF SURFACE TURBULENTFLUX MODELS. Bulk flux algorithms link turbu-lent fluxes to mean values of surface temperature,wind, and surface air temperature and humidity, aver-aged over the time scale of multiple turbulent eddies:

t = raCDU2, (1)

LH = raLvCEU(qs – qa), (2)

SH = racpCHU (qs – qa), (3)

where t, LH, and SH are, respectively, the wind stress,latent heat flux, and sensible heat flux; ra is the den-sity of air; Lv is the latent heat of vaporization; cp is thespecific heat of air at constant pressure, U is the near-surface wind speed relative to the surface currentspeed; q is specific humidity; q is the potential tem-perature; the subscripts s and a are, respectively, thevalues at the surface and the near-surface atmosphere(usually the standard height of 10 m above sea level);and CD, CE, and CH are, respectively, the turbulent ex-change coefficients for momentum, moisture, and heat.

A variety of different bulk flux algorithms are pres-ently used, with the most recently developed research-

quality algorithms showing fairly good agreementwith each other and with observations in conditionsof moderate wind speeds and neutral or unstable con-ditions (e.g., Smith 1988; Fairall et al. 2003; Claysonet al. 1996; Zeng et al. 1998; Bourassa et al. 1999).While these algorithms all use (1)–(3) to calculate theturbulent fluxes, they differ in their parameterizationof the exchange coefficients (including treatment ofsurface roughness length), salinity effect on sea sur-face humidity, and treatment of free convective con-ditions and surface layer gustiness.

The in situ cases with direct turbulent measure-ments are being used to evaluate both the bulk calcu-lations of turbulent fluxes and specific parameter-izations within the bulk models. These in situ casesare characterized by predominantly neutral to un-stable stratification in moderate wind speeds. Thereare several cases with high wind speeds, notably thecases at high latitudes. There are no direct flux mea-surements available under precipitating or sea sprayconditions.

Brunke et al. (2002; 2003) have conducted a pre-liminary comparison of a variety of bulk flux algo-rithms with ship observations. Presented here aresome results from four bulk flux models: TOGACOARE version 3.0 (COARE 3.0), and those used toproduce the satellite-derived datasets HOAPS,GSSTF-2, and J-OFURO. Figure 2 presents an evalu-ation of the modeled latent heat (LH) fluxes and di-rect flux measurements taken during the NOAA ETLand French deployments. The COARE 3.0 algorithmhas the lowest bias. The biases for each algorithm werealso found to vary when compared against differentin situ cases, as well as with wind speed. Large biasesalso exist in sensible heat flux and wind stress; windstress biases for these four algorithms at low windspeeds can be over 30%, although the absolute error

FIG. 2. Evaluation of modeled latent heat fluxes and di-rect flux measurements taken during the NOAA ETLand French deployments. The following bulk flux mod-els are compared (see text): COARE 3.0, HOAPS,GSSTF-2, and J-OFURO.

Page 8: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

416 MARCH 2004|

(N m-2) at low wind speeds is small . As a result ofthese comparisons, HOAPS and GSSTF-2 plan to usethe COARE 3.0 model in the next versions of thesedatasets.

This comparison indicates that within the accu-racy of the direct turbulent flux measurements, thebulk flux models used in preparation of the satellitedatasets could be improved. While each of these bulkflux models has been evaluated, using a limitedamount of in situ data, the more extensive evaluationover a variety of regimes has revealed some signifi-cant biases. The COARE 3.0 bulk model, which hasbeen developed primarily for interpreting in situ tur-bulent flux measurements, and has hitherto not beenused in the preparation of satellite datasets, could beused to improve the accuracy of satellite-derived sur-face turbulent fluxes for some conditions. Additionaldevelopments in surface flux models (e.g., Claysonet al. 1996; Bourassa et al. 1999) could also be usedto improve the calculation of surface turbulent fluxesfrom satellite observations. SEAFLUX will use infor-mation from this comparison to guide developmentof a more accurate model that applies to a wider rangeof conditions.

The following are outstanding issues in bulk al-gorithms: conditions of light wind and stable strati-fication, treatment of sea state (swell and directionaleffects), appropriate averaging scales, and parameter-ization of mesoscale gustiness.

INPUT STATE VARIABLES. To calculate sur-face turbulent fluxes from the bulk algorithms, thefollowing input variables are required: sea surfacetemperature, surface wind, and surface air tempera-ture and specific humidity. Issues in determiningeach of these variables from the satellite are describedbelow.

Sea surface temperature. Historically, the bulk SST hasbeen used to compute the air–sea heat fluxes becauseit has only been more recently possible to measure theradiative skin SST. Because of its more direct relation-ship to the surface turbulent fluxes, in SEAFLUX weplan to use the skin SST in determining the heatfluxes. Operational satellite SST products (e.g.,Reynolds and Smith 1995) are not suitable forSEAFLUX because they provide estimates of bulkSSTs, and not the skin value that determines the sur-face fluxes. Moreover, because operational SST prod-ucts rely solely on polar-orbiting satellite infraredmeasurements, they only provide values once everyfew days (much less frequently in the Tropics andmidlatitude storm track regions) under cloud-free

conditions and do not resolve systematic diurnalvariations. Thus, although the spatial sampling can,in principle, be performed at 4 km, the product avail-able represents bulk SST values at 5-day intervals,smoothed to about 100 km spatially. To determine thebulk turbulent fluxes and the upwelling surfacelongwave radiative flux, we need to create a new skinSST dataset that resolves the diurnal cycle under bothclear and cloudy conditions. SST products based onAVHRR cannot resolve diurnal variations or providevalues in cloudy conditions. While the operationalgeostationary satellites offer diurnal sampling, itcomes with a significant reduction in spatial resolu-tion, and the 12-mm channel for “split window” re-trievals is not available on all satellites.

Microwave radiometers that include lower-frequency channels near 6–10 GHz can be used to in-fer a temperature that is slightly deeper than the skinvalue, because the penetration depth into water atthese frequencies is 1–2 mm, under almost all cloudyconditions (as well as clear). The only low-frequencymicrowave data currently available comes from theTMI; although the orbit of this satellite drifts in localtime to provide a statistical sample of the diurnal SSTvariations, it does not provide high time resolutionmeasurements. Moreover, its coverage is limited tothe Tropics and subtropics. Nevertheless, these datawill be used to verify other methods for determininga high time resolution skin SST product under all me-teorological conditions. The advanced microwavescanning radiometer (AMSR) instrument on NASA’sEarth Observing System (EOS) Aqua (launched 4 May2002), and the AMSR EOS (AMSR-E) on the Japa-nese Advanced EOS (ADEOS) will allow for globalmicrowave measurements of SSTs.

Several approaches have been proposed for deter-mining high-resolution skin SST values. Fairall et al.(1996) and Zeng et al. (1999) use a diurnally varyingbulk skin correction to determine skin SST valuesfrom the operational bulk SST products. This com-prises separate models for the cool skin and the di-urnal warming above the depth at which the bulkmeasurement is made. The ISCCP (Rossow andGarder 1993) determines SSTs using clear-sky infra-red radiances and determines a diurnal cycle ampli-tude from space–time-composited clear-sky SST val-ues. Clayson and Curry (1996) developed a methodfor determining the diurnal cycle of skin SSTs thatuses a cosine-shaped diurnal cycle of skin SSTs(which is a model-derived function of surface windspeed, solar insolation, and precipitation) superim-posed on a time series of predawn skin SST valuesderived from available (cloud free) infrared retrievals.

Page 9: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

417MARCH 2004AMERICAN METEOROLOGICAL SOCIETY |

The SEAFLUX intercomparison will evaluate thesemethods for determining high-resolution global skinSST products. Some preliminary results from the SSTintercomparison are shown in Fig. 3 for shipboardobservations obtained from the R/V Ron Brown in thewestern tropical Pacific during July 1999. During theperiod spanned by days 182–197, the ship was nearlystationary in the vicinity of the island of Nauru at0.5oS, 167oE. From the period spanned by days 165–200, the ship traveled from Darwin, Australia, to theKwajalein Atoll by way of Nauru. The in situ SSTmeasurements were obtained using the ETL SeaSnake, which consists of a floating thermistor thatnominally measures SST at a depth of 5 cm. For aperiod of 3 days (days 182–184), the R/V Ron Brownwas collocated with the R/V Mirai, making radiomet-

ric skin SST measurements using the Marine Atmo-spheric Emitted Radiance Interferometer (MAERI;Smith et al. 1996). Here, we use Castro’s (2001) bulkskin model with the Sea Snake bulk SST, which pro-vides skin SST values that are very close to the MAERIobservations. Because radiometric skin SSTs are rela-tively rare and proprietary, here we use Castro’s bulkskin SST model with the Sea Snake data to provide arobust and inexpensive proxy for radiometric skinSST measurements. A comparison of the Sea Snakemeasurement and the modeled skin SST value isshown in Fig. 3a.

Figure 3a also shows the TMI (passive microwave)measurements of SSTs produced by F. Wentz of Re-mote Sensing Systems (Wentz et al. 2000). The TMIvalues are biased high, relative to the skin SST values

The temperature at theinterface between the atmosphereand ocean is called the skin seasurface temperature. It is thisinterfacial temperature thatshould be used to calculate thesurface sensible and latent heatfluxes and the upwelling longwaveflux. The skin temperature cannotbe measured directly usingpresent technology, so remoteinfrared thermometers areemployed to sense the radiativeskin SST. The so-called sea surfacetemperatures are most commonlymeasured from ships with ther-mometers by sampling water at adepth up to 5 m from enginewater intake or buckets, or frombuoys or moorings that measuretemperature at a depth from 0.5to 1.5 m. These measurementsare referred to as bulk sea surfacetemperatures and are typicallycharacteristic of the temperatureof the ocean mixed layer sometens of meters deep. Observationsshow that the skin temperature isinvariably a few tenths of a degreecooler than the water a fewmillimeters below the surface,even during periods of weak windsand strong insolation.

To explain the cool skin, weexamine the energy balance of amillimeter-thick layer at the oceansurface. While virtually all of the

shortwave radiation is absorbed inthe ocean mixed layer, less than10% is absorbed in the uppermillimeter. Because the surfacelatent and sensible heat fluxes andthe net longwave radiation fluxesare typically negative (cool theocean surface), there is a net heatloss even in this millimeter-thickskin layer, although the oceanmixed layer may be heating due tosolar radiation. The net heat lossin the thin surface layer requires aflux of heat from the upper ocean.On both sides of the interface, theatmosphere and ocean aretypically in turbulent motion.However, upon approaching theinterface, turbulence is suppressedby the large density contrastbetween the air and ocean, andthe interface is a strong barrier tothe turbulent transport betweenthe ocean and atmosphere.Therefore, on both sides of theinterface the required heattransfer is accomplished bymolecular conduction. Theconsequence of the large heat lossat the surface is that the tempera-ture gradient just below thesurface is large and negative in theupward direction. Such a tem-perature gradient allows heat fluxfrom the ocean interior to balancethe surface loss. This results in acool skin that is a few tenths of a

degree cooler than the oceantemperature 1 mm below thesurface.

The temperature drop acrossthe viscous sublayer just below theocean surface, DDDDDT, is the bulk skintemperature difference. Typicalnighttime values of DDDDDT are 0.3°C,although values may exceed 1°Cunder some extreme conditions.During the daytime there issignificant variability in DDDDDT thatdepends on the amount of solarinsolation, ocean turbidity, and themagnitude of the wind. It ispossible for the skin SST tobecome warmer than the bulkSST at a 3–5-m depth, due todiurnal warming of the upper fewmeters above the diurnal ther-mocline. Relative to the tempera-ture at a few centimeters depth,the skin SST is still “cool” byabout 0.3 K. While such smallvalues of DDDDDT may seem insignifi-cant, use of the bulk temperatureinstead of the skin temperature tocalculate the surface sensible andlatent heat fluxes from (2) and (3)can result in systematic errors inthe computed fluxes that exceed10%. Because of the Clausius–Clapeyron relationship, smallerrors in surface temperatureresult in larger errors in the latentheat flux, particularly when thesurface temperature is high.

SKIN SST

Page 10: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

418 MARCH 2004|

(0.63oC), showing less bias when compared with thedirect Sea Snake measurements (0.38oC). Random er-rors in the TMI are seen, with root-mean-square er-rors relative to the Sea Snake temperature of 0.82oC,presumably reflecting deficiencies in the TMI calibra-tion and/or the retrieval method.

Figure 3b compares the Sea Snake SST and themodeled skin SST with the Naval Oceanographic Of-fice (NAVO) nonlinear SST bulk algorithm (May etal. 1998) applied to both 1-km AVHRR high-resolu-tion picture transmission (HRPT) data and 4-kmGAC data. Some of the retrieved values are greaterthan 1o cooler than the modeled skin SST values, sug-gesting errors in cloud clearing. The pixels that do notappear contaminated by clouds tend to be slightlywarmer than both the bulk and the modeled skin val-ues, but are generally within 1oC.

Neither the TMI nor NAVO SST products resolvethe diurnal variation of skin SST. Two methods havebeen proposed to determine the diurnal cycle of skinSST from the satellite. The first method (Fig. 3c) isthat of the ISCCP (Rossow and Garder 1993), wherethe diurnal cycle is determined from a composite ofclear-sky pixels (AVHRR) in a 280-km grid cell overa 15-day period. The ISCCP method has a bias of+0.02oC and a root-mean-square error of 2.02oC. Be-cause the ISCCP method is determined from clear-sky pixels, the diurnal amplitude is significantlygreater than the observed amplitude. The secondmethod (Fig. 3d; Clayson and Curry 1996) uses amodel-derived diurnal cycle amplitude that requiresas input the peak solar insolation, surface wind speed,and daily averaged precipitation. The Clayson–Currymethod has a bias of +0.33oC and a root-mean-square

FIG. 3. Evaluation of satellite methods to determine SSTs. Observations are obtained from the NOAA ETL SeaSnake onboard the R/V Ron Brown during the period 14 Jun 1999–19 Jul 1999: (a) Sea Snake SST measurements(dash), modeled skin SSTs (solid), TMI values (solid diamonds ); (b) Sea Snake SST measurements (dash), mod-eled skin SSTs (solid), NAVO skin SSTs derived from AVHRR (solid circles); (c) modeled skin SST, (solid), ISCCPskin SSTs derived from AVHRR (dash ); and (d) modeled skin SSTs (solid), Clayson–Curry skin SSTs (dash).

Page 11: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

419MARCH 2004AMERICAN METEOROLOGICAL SOCIETY |

error of 0.62oC. The amplitude of the diurnal cycle isvery close to the observed amplitude. The primarysource of error in the Clayson–Curry method is theselection of the predawn skin SSTs. This is presentlyaccomplished by using the NOAA operational bulkSST product (with 5-day resolution), and by apply-ing a bulk skin correction. Note that the NOAA op-erational SST product is derived from the NAVO skinSST values (Fig. 3b), which are regressed against buoydata to determine a bulk SST and are then averaged.

While further comparisons are needed, these re-sults suggest that

• skin SST models are now sufficiently accurate(Castro 2001) so that floating thermistors, such asthe Sea Snake, can be used to determine skin SSTsfor satellite validation;

• methods to infer the diurnal cycle of skin SST ap-pear promising, particularly in the Tropics wherethe diurnal cycle of skin SST is the greatest; and

• a combination of the passive microwave, skin SSTmodel, infrared, and diurnal cycles should be suf-ficient to produce a SST product with the desiredaccuracy and resolution.

Winds. Several global wind datasets are available, in-cluding satellite, ship, and NWP products. Passivemicrowave radiometers provide the foundation forseveral global datasets of wind speed (e.g., Wentz1997). Scatterometers, which measure backscatter

from the sea surface to provide global near-surfacewind speeds and directions, provide the most prom-ising data stream for vector winds (e.g., Pegion et al.2000). Algorithms for determining winds fromscatterometers are highly empirical and realistic er-ror models are needed (Bourassa et al. 2003).Parametric issues in scatterometer wind determina-tions include the following: retrievals at high and lowwind speed extremes, stratification extremes (stable,unstable), retrievals in precipitating conditions, andeffects of surface currents and waves. More researchis needed to determine the time and space resolutionfor which vector winds can be determined from mul-tiple scatterometers (e.g., Schlax et al. 2001). Thescatterometer community is conducting some in situevaluations of wind products (e.g., Verschell et al.1999). It is not straightforward to evaluate vectorwind products using in situ measurements (Freilichand Dunbar 1999; Bourassa et al. 2003). Hence, thefocus of our intercomparison will be on scalar windsthat are used within the parameterization schemes forturbulent heat fluxes; we are coordinating closelywith other groups that are evaluating vector windsdetermined from satellites.

To demonstrate the potential utility of scat-terometer wind observations relative to NWP prod-ucts, Fig. 4 illustrates a case in the Gulf of Mexico on20 September 1999, whereby the high-resolutionNCEP Eta Model analyzed a tropical wave. Analysisof SeaWinds scatterometer data (following Pegion

FIG. 4. Comparison of surface wind stress analyzed in the Gulf of Mexico at 0000 UTC 20 Sep 1999. Stresses arecalculated with a wind speed–dependent drag coefficient. (a) The 22-km Eta NWP analysis shows a tropical wave.(b) The scatterometer product, on a coarser 0.5° grid, shows a much deeper and more detailed Tropical StormHarvey soon after the National Hurricane Center classified Harvey as reaching tropical storm strength.

Page 12: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

420 MARCH 2004|

et al. 2000) for a 3-h period, centered on the sametime, shows a closed tropical storm with surface windstress reaching values an order of magnitude higherthan the Eta analysis. While the Eta Model analysesare quite effective in this region for episodic forcingrelated to rapidly moving cold fronts, this example istypical of tropical cyclones. Despite the fine grid spac-ing in many NWP models, these models lack the reso-lution that can be achieved from satellite observationswith similar in-swath grid spacing, even after the sat-ellite data are regridded to a coarser regular grid. Notethat this high-quality satellite product is currentlyonly achievable regionally (due to the satellite sam-pling pattern), and that additional satellite coverageis needed to achieve this quality and resolution in glo-bal products.

Daily global gridded surface wind parameters, in-cluding wind vectors, stress, curl, and divergence,have been computed since August 1991 from threesatellite microwave scatterometers (Bentamy et al.2003): the Active Microwave Instrument (AMI)[onboard the European Remote Sensing Satellites(ERS-1 and ERS-2)], and NSCAT (onboardADEOS1), which have a spatial resolution of 50 kmover a swath of 500 km, and twice 600 km, respec-tively. These gridded winds have been used exten-sively in global wind studies (Bentamy et al. 1998),and in ocean model forcing (Grima et al. 1999;Quilfen et al. 2000). The root-mean-square differencebetween satellite and buoy wind estimates are1.16 m s-1 for wind speed and 30° for wind direction(Bentamy et al. 1998). Recently, synthetic apertureradars (SARs) have been used to determine oceansurface vector winds (e.g., Perrie et al. 2003). Theresolution of the SAR (2 km) is significantly higherthan that for the scatterometers, but the narrowswath width and, hence, low spatial and temporalcoverage reduces the utility of this dataset forSEAFLUX. For scalar winds, the combination ofscatterometer and passive microwave measurementsshould provide the needed time/space resolution forSEAFLUX.

Surface air temperature and humidity. Bulk formulas forsurface turbulent fluxes require information about thedifference between the ocean surface skin tempera-ture and humidity (a function of skin surface tempera-ture and salinity), and the temperature (Ta) and hu-midity (qa) of the near-surface atmosphere. While theskin temperature can be almost directly sensed fromsatellites, the properties of the near-surface atmo-sphere are difficult to determine. Satellite tempera-ture and humidity sounders provide profiles of these

quantities in the atmosphere, but do not directly re-solve the planetary boundary layer.

To determine qa from the satellite, all publishedmethods make use of passive microwave instruments.Early methods used column precipitable water.Recent methods that separate upper- and lower-tropospheric precipitable water using passive micro-wave observations include Schulz (1997), Chou et al.(1997), Lin et al. (2001), and Jones et al. (1999).Schlussel et al. (1995; which is used in the J-OFUROdataset) directly determine qa from the passivemicrowave observations. The peaks of the humidityweighting functions in current satellite profilingsystems are at levels generally much higher than thesea surface. Using NOAA radiosonde data and amicrowave radiative transfer model, it was deter-mined that estimated qa values from microwave hu-midity sounders [i.e., Advanced Microwave Sound-ing Unit (AMSU)-B and Humidity Sounder forBrazil (HSB)] have larger errors (about 1.7 g kg-1)than those (~1.1 g kg-1) retrieved from integratedwater vapor measurements (i.e., SSM/I and TMI)whose weighting functions have a peak at the surface.However, the combination of the Special SensorMicrowave Water Vapor Sounder (SSM/T2) with in-tegrated water vapor measurements from SSM/I pro-vided better results than from either instrumentwhen used alone.

No space-based instrument can give accurate, di-rect estimates of air temperature in the lower atmo-sphere. Early methods to determine Ta from satelliteobservations used a specified value of the relative hu-midity with the value of qa (e.g., Kubota andShikauchi 1995; Konda et al. 1996). Kubota andMitsumori (1997) obtained better sensible heat fluxesby using a specified Bowen ratio to determine thesensible heat flux directly from the latent heat flux.Other methods have related Ta to precipitable waterand U (Jones et al. 1999), or to U and qa (Konda etal. 1996). These methods do not provide the spatialand temporal resolution desired by SEAFLUX. Sev-eral recent efforts have been successful at determin-ing high-resolution values of either Ta or Ta – Ts forspecific regions: in the Tropics Clayson and Curry(1996) found a robust statistical relationship betweenTa – Ts and a satellite cloud classification scheme thatconsidered cloud-top height, precipitation, and dayversus night; and Bourras et al. (2002b) used a simplehorizontal temperature advection model to derive airtemperature and sensible heat flux fields from satel-lite observations of wind and SST at the scale of a1000-km region. An additional method using up-stream values of Ts to estimate Ta is being explored.

Page 13: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

421MARCH 2004AMERICAN METEOROLOGICAL SOCIETY |

The value of existing satellite temperature profile dataon estimating Ta is also being investigated. UntilSEAFLUX, there was little motivation to determinehigh-resolution values of Ta, and, hence, the status ofthis investigation is preliminary.

While there are some promising new ideas for im-proving the satellite-derived values of qa and Ta, at thispoint (especially for Ta) NWP analyses presently pro-duce better values of Ta and qa at the desired resolu-tion. As a result, Chou et al. (2003) have directly in-corporated NWP values of Ta in their GSSTF satelliteflux product. While this combination of satellite andNWP analyses arguably provides the most accuraterepresentation of surface fluxes presently available,such a hybrid flux dataset cannot be used to evaluateNWP analyses because it is not an independentdataset.

Substantially improved profiling capability for at-mospheric temperature and humidity (Zhou et al.2002) will be available from the National Polar-Orbiting Operational Environmental Satellite System(NPOESS).

SUMMARY AND CONCLUSIONS. TheSEAFLUX strategy to assess satellite flux products,and to improve the accuracy and resolution ofsatellite-derived sensible and latent heat fluxes, hasbeen described. Our analysis shows that at present,zonal and monthly averaged flux values have signifi-cant uncertainties, based on the comparison ofclimatologies determined from ships, numericalweather prediction analyses, and satellite products.Uncertainties are even greater when these data areconsidered at higher spatial and temporal resolu-tions, as required by applications such as oceanmodeling.

Evaluation of bulk aerodynamic flux models sug-gests that those used to develop global satellite fluxproducts are not the most accurate algorithms avail-able. Outstanding issues in bulk flux models are asfollows: conditions of light wind and stable stratifi-cation, treatment of sea state (swell and directional ef-fects), appropriate averaging scales, and parameter-ization of mesoscale gustiness.

Our preliminary analysis of available methods fordetermining skin sea surface temperature indicatesthat skin SST models are now sufficiently accurate sothat simple floating thermistors can be used to deter-mine skin SSTs for satellite validation. Methods to in-fer the diurnal cycle of skin SST are promising, par-ticularly in the Tropics where the diurnal cycle of skinSST is the greatest. A combination of the passive mi-crowave, skin SST model, infrared, and diurnal cycles

should be sufficient to produce a skin SST productwith the desired accuracy and resolution.

Surface winds from scatterometers provide moredetailed analysis of storms than is presently feasibleusing numerical weather prediction models. Paramet-ric issues in scatterometer wind determinations in-clude retrievals at high and low wind speed extremes,stratification extremes (stable, unstable), retrievals inprecipitating conditions, and effects of surface cur-rents and waves. More research is needed to deter-mine the time and space resolution for which vectorwinds can be reliably determined from multiplescatterometers. Merging scatterometer products withpassive microwave winds should provide the neededtime/space resolution for SEAFLUX.

At relatively low spatial and temporal resolutions,surface air specific humidity has been reliably deter-mined by passive microwave imagers. Increasing theresolution increases the root-mean-square error of thevalues. Satellite determination of surface air tempera-ture, or the difference between the air and surfacetemperatures, have been inferred by indirect meth-ods, some of which have been shown to be successfulin regional studies. Use of profilers, in conjunctionwith other techniques, is being investigated to deter-mine surface air temperature and humidities; the nextgeneration of operational satellites is expected to in-clude improved profiling capability. Because the root-mean-square error in satellite-derived bulk variablesaffects the accuracy of the retrieved fluxes, Bourras etal. (2002a) developed a direct relationship betweensurface latent heat flux and satellite observations us-ing a neural network approach. The neural networkmethod relates microwave brightness temperaturesand SST values from infrared sensors to surface latentheat fluxes. The in situ dataset assembled bySEAFLUX could be used for an optimal developmentof such a neural network.

The currently available global satellite ocean sur-face turbulent flux datasets (HOAPS, GSSTF, and J-OFURO) are based on passive satellite microwavemeasurements (SSM/I), and are supplemented bySSTs derived from infrared observations (AVHRR).Bentamy et al. (2003) have produced a global datasetfor 1 yr, incorporating scatterometer data (along withSSM/I and AVHRR data). To enhance the accuracyand the time/space resolution of turbulent flux fieldsover global oceans, the merging of multiple satellitedatasets is desired.

Even if the accuracy goal of 1o spatial resolution,3–6 h time resolution, and 5 W m-2 for individualcomponents of the surface heat budget, is not achievedwith currently available measurements, the results of

Page 14: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

422 MARCH 2004|

the intercomparison project will substantially im-prove global flux products, and the development ofthese products will highlight the obstacles to achiev-ing the required accuracy. In any case, the resultsshould be useful for determining the dominant scalesof variability and identifying the responsible pro-cesses. Carefully characterized errors in the fluxesallow an imperfect dataset to be used at some level forthe target applications. Incorporation of some vari-ables from NWP analyses (e.g., surface air tempera-ture and possibly surface air specific humidity) canimprove the accuracy of the satellite flux products,although such a hybrid flux product cannot be usedto evaluate NWP models.

While SEAFLUX has assembled a substantial li-brary of in situ datasets, we are seeking additional re-search-quality datasets that include direct turbulentflux measurements, wave information, or radiomet-ric skin SST measurements. We anticipate that im-proved surface turbulent flux products will continueto be produced as a result of SEAFLUX. We are veryinterested in obtaining any feedback on the successor failures of applications using satellite- or NWP-derived ocean surface turbulent flux products

ACKNOWLEDGMENTS. Coordination of theSEAFLUX project has been funded by a grant from NASAEOS IDS (PI: P. Webster).

REFERENCESBentamy, A., N. Grima, and Y. Quilfen, 1998: Valida-

tion of the gridded weekly and monthly wind fieldscalculated from ERS-1 scatterometer wind observa-tions. Global Atmos. Ocean Syst., 6, 373–396.

——, K. Katsaros, A. Mestas-Nuñez, W. Drennan, E.Forde, and H. Roquet, 2003: Satellite estimates ofwind speed and latent heat flux over the globaloceans. J. Climate, 16, 637–656.

Bourassa, M. A., D. G. Vincent, and W. L. Wood, 1999:A flux parameterization including the effects of cap-illary waves and sea state. J. Atmos. Sci., 56, 1123–1139.

——, D. M. Legler, J. J. O’Brien, and S. R. Smith, 2003:SeaWinds validation with research vessels. J.Geophys. Res., 108, 3019, doi: 10.1029/2001JC001028.

Bourras, D., L. Eymard, and W. T. Liu, 2002a: A neuralnetwork to estimate the latent heat flux over oceansfrom satellite observations. Int. J. Remote Sens., 23,2405–2423.

——, ——, ——, and H. Dupuis, 2002b: An integrated ap-proach to estimate instantaneous near-surface airtemperature and sensible heat flux fields during the

SEMAPHORE experiment. J. Appl. Meteor., 41, 241–252.

Bruemmer, B., 1993: ARKTIS 1993: Report on the fieldphase with examples of measurements. Ber. ZentrumMeeres Klimaforschung Reihe A, 11.

Brunke, M. A., X. Zeng, and S. Anderson, 2002: Uncer-tainties in sea surface turbulent flux algorithms anddata sets. J. Geophys. Res., 107, 3141, doi: 10.1029/2001JC000992.

——, C. W. Fairall, and X. Zeng, 2003: Which bulk aero-dynamic algorithms are least problematic in comput-ing ocean surface turbulent fluxes? J. Climate, 16,619–635.

Castro, S. L., 2001: Further refinements to models for thebulk-skin sea surface temperature difference. Ph. D.thesis, University of Colorado, 187 pp.

Chou, S. H., C. L. Shie, R. M. Atlas, and J. Ardizzone,1997: Air-sea fluxes retrieved from special sensormicrowave imager data. J. Geophys. Res., 102, 12 705–12 726.

——, ——, ——, ——, and E. Nelkin, 2001: A multi-yeardata set of SSM/I-derived global ocean surface tur-bulent fluxes. Preprints, 11th Conf. on Interaction ofthe Sea and Atmosphere, San Diego, CA, Amer. Me-teor. Soc., 131–134.

——, E. Nelkin, J. Ardizzone, R. M. Atlas, and C.-L. Shie,2003: Version 2 Goddard Satellite-Based SurfaceTurbulent Fluxes (GSSTF-2). Preprints, 12th Conf. onInteractions of the Sea and Atmosphere, Long Beach,CA, Amer. Meteor. Soc., 3.6.

Clayson, C. A., and J. A. Curry, 1996: Determination ofsurface turbulent fluxes for TOGA COARE: Com-parison of satellite retrievals and in situ measure-ments. J. Geophys. Res., 101, 28 503–28 513.

——, C. W. Fairall, and J. A. Curry, 1996: Evaluation ofturbulent fluxes at the ocean surface using surfacerenewal theory. J. Geophys. Res., 101, 28 515–28 528.

Curry, J. A., C. A. Clayson, W. B. Rossow, R. R. Reeder,Y. C. Zhang, P. J. Webster, G. Liu, and R. S. Sheu,1999: High-resolution satellite-derived dataset of thesurface fluxes of heat, freshwater and momentum forthe TOGA COARE IOP. Bull. Amer. Meteor. Soc., 80,2059–2080.

DaSilva, A. M., C. C. Young, and S. Levitus, 1994: Algo-rithms and Procedures. Vol. 1. Atlas of Surface Ma-rine Data 1994. NOAA Atlas NESDIS 6, 83 pp.[Available online at http://ingrid.ldeo.columbia.edu/S O U R C E S / . D A S I L V A / . S M D 9 4 / . d a t a s e t _documentation.html.]

Ebert, E. E., M. J. Manton, P. A. Arkin, R. J. Allam, G. E.Holpin, and A. Gruber, 1996: Results from the GPCPAlgorithm Intercomparison Programme. Bull. Amer.Meteor. Soc., 77, 2875–2887.

Page 15: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

423MARCH 2004AMERICAN METEOROLOGICAL SOCIETY |

Edson, J. B., A. A. Hinton, K. E. Prada, J. E. Hare, andC. W. Fairall, 1998: Direct covariance flux estimatesfrom moving platforms at sea. J. Atmos. OceanicTechnol., 15, 547–562.

Emanuel, K., 2001: Contribution of tropical cyclones tomeridional heat transport by the oceans. J. Geophys.Res., 106, 14 771–14 781.

Emery, W. J., S. Castro, G. A. Wick, P. Schluessel, andC. Donlon, 2001: Estimating sea surface temperaturefrom infrared satellite and in situ temperature data.Bull. Amer. Meteor. Soc., 82, 2773–2785.

Fairall, C. W., E. F. Bradley, D. P. Rogers, J. B. Edson,and G. S. Young, 1996: Bulk parameterization of air-sea fluxes in TOGA COARE. J. Geophys. Res., 101,3747–3767.

——, A. B. White, J. B. Edson, and J. E. Hare, 1997:Integrated shipboard measurements of the marineboundary layer. J. Atmos. Oceanic Technol., 14, 338–359.

Fasullo, J. T., and P. J. Webster, 2000: Structure of theocean-atmosphere system in the tropical western Pa-cific during strong westerly wind bursts. Quart. J.Roy. Meteor. Soc., 126, 899–924.

Freilich, M. H., and R. S. Dunbar, 1999: The accuracy ofthe NSCAT 1 vector winds: Comparisons with Na-tional Data Buoy Center buoys. J. Geophys. Res., 104,11 231–11 246.

Godfrey, J. S., E. F. Bradley, P. A. Coppin, L. Pender, T.J. McDougall, E. W. Sculz, and I. Helmond, 1999:Measurements of upper ocean heat and freshwaterbudgets near a drifting buoy in the equatorial Indianocean. J. Geophys. Res., 104, 13 269–13 302.

Grima, N., A. Bentamy, K. Katsaros, and Y. Quilfen,1999: Sensitivity of an oceanic general circulationmodel forced by satellite wind stress fields. J. Geophys.Res., 104, 7967–7989.

Gupta, S. K., N. A. Ritchey, A. C. Wilber, C. H. Whitlock,G. G. Gibson, and P. W. Stackhouse, 1999: A clima-tology of surface radiation budget derived from sat-ellite data. J. Climate, 12, 2691–2710.

Hayes, S. P., M. J. McPhaden, and A. Leetmaa, 1989:Quasi real-time simulation of the tropical PacificOcean. J. Geophys. Res., 94, 2147–2157.

Jones, C., P. Peterson, and C. Gautier, 1999: A newmethod for deriving oean surface specific humidityand air temperature. An artificial neural network. J.Appl. Meteor., 38, 1229–1245.

Josse, P., G. Caniaux, and S. Planton, 1999:Intercomparison of ocean-atmosphere fluxes fromoceanic and atmospheric forced and coupled mesos-cale simulations. Ann. Geophys., 17, 566–576.

Konda, M., N. Imasato, and A. Shibata, 1996: Analysisof the global relationship of biennial variation of sea

surface temperature and air-sea heat flux using sat-ellite data. J. Oceanogr., 52, 717–746.

Kubota, M., and A. Shikauchi, 1995: Air temperature atocean surface derived from surface-level humidity.J. Oceanogr., 51, 619–634.

——, and S. Mitsumori, 1997: Sensible heat flux esti-mated by using satellite data over the North Pacific.Remote Sensing of the Subtropical Ocean, C. T. Liu,Ed., Elsevier, 127–136.

——, N. Iwasaka, S. Kizu, and M. Knoda, 2002: JapaneseOcean Flux Data Sets with Use of Remote SensingObservations (J-OFURO). J. Oceanogr., 58, 213–215.

——, A. Kano, and H. Tomita, 2003: Intercomparisonof various surface latent heat flux fields. J. Climate,16, 670–678.

Lin, B., P. Minnis, A. Fan, T. Charlock, D. Young, andY. Hu, 2001: The surface and TOA heat budget overtropical oceans observed by TRMM satellite. Pre-prints, 11th Conf. on Interaction of Sea and Atmo-sphere, San Diego, CA, Amer. Meteor. Soc., 127–129.

Marshall, J., and Coauthors, 1998: The Labrador SeaDeep Convection Experiment. Bull. Amer. Meteor.Soc., 79, 2033–2058.

May, D. A., M. M. Parmeter, D. S. Olszewski, and B. D.McKenzie, 1998: Operational processing of satellitesea surface temperature retrievals at the NavalOceanographic Office. Bull. Amer. Meteor. Soc., 79,397–407.

McGillis, W. R., J. B. Edson, J. E. Hare, and C. W. Fairall,2001: Direct covariance air-sea CO2 fluxes. J.Geophys. Res., 106, 16 729–16 745.

Moore, G. W. K., and I. A. Renfrew, 2002: An assessmentof the surface turbulent heat fluxes from the NCEPreanalysis over western boundary currents, J. Cli-mate, 15, 2020–2037.

Nardelli, B. B., and E. Salusti, 2000: On dense water for-mation criteria and their application to the Mediter-ranean Sea. Deep Sea Res., 47, 193–221.

Pegion, P. J., M. A. Bourassa, D. M. Legler, and J. J.O’Brien, 2000: Objectively derived daily “winds”from satellite scatterometer data. Mon. Wea. Rev.,128, 3150–3168

Perrie, W., E. Dunlap, P. W. Vachon, B. Toulany, R.Anderson, and M. Dowd, 2003: Marine wind analy-sis from remotely sensed measurements. Can. Re-mote Sens., 28, 450–465.

Quilfen, Y., A. Bentamy, P. Delecluse, K. Katsaros, andN. Grima, 2000: Prediction of sea level anomalies us-ing ocean circulation model forced by scatterometerwind and validation using TOPEX/Poseidon data.IEEE Trans. Geosci. Remote Sens., 38, 1871–1884.

Renfrew, I. A., G. W. K. Moore, P. S. Guest, and K.Bumke, 2002: A comparison of surface layer and sur-

Page 16: SEAFLUX - Judith Curry · 2006-10-20 · region (Curry et al. 1999), the tropical Pacific Ocean (Jones et al. 1999), the global tropical oceans (Lin et al. 2001; Schlussel and Albert

424 MARCH 2004|

face turbulent flux observations over the LabradorSea with ECMWF analyses and NCEP reanalyses. J.Phys. Oceanogr., 32, 383–400.

Reynolds, W. R., and T. M. Smith, 1995: A high-resolu-tion global sea surface temperature climatology. J.Climate, 8, 1571–1583.

Rossow, W. B., and L. C. Garder, 1993: Cloud detectionusing satellite measurements of infrared and visibleradiances for ISCCP. J. Climate, 6, 2341–2369.

——, and R. A. Schiffer, 1999: Advances in understand-ing clouds from ISCCP. Bull. Amer. Meteor. Soc., 80,2261–2287.

Schlax, M. G., D. B. Chelton, and M. H. Freilich, 2001:Sampling errors in wind fields constructed fromsingle and tandem scatterometer datasets. J. Atmos.Oceanic Technol., 18, 1014–1036.

Schlüssel, P., and A. Albert, 2001: Latent heat flux at thesea surface retrieved from combined TMI and VIRSmeasurements of TRMM. Int. J. Remote Sens., 22,2861–2861

——, L. Schanz, and G. Englisch, 1995: Water vapor inthe atmospheric boundary layer over oceans fromSSM/I and AVHRR measurements. Adv. Space Res.,16, 107–116.

Schulz, J., J. Meywerk, S. Ewald, and P. Schlüssel, 1997:Evaluation of satellite-derived latent heat fluxes. J.Climate, 10, 2782–2795.

Smith, S. D., 1988: Coefficients for sea surface windstress, heat flux, and wind profiles as a function ofwind speed and temperature. J. Geophys. Res., 93, 15467–15 472.

Smith, S. R., C. Harvey, and D. M. Legler, 1996: Hand-book of quality control procedures and methods forsurface meteorology data. WOCE Data AssemblyCenter for Surface Meteorology, Center for OceanAtmospheric Prediction Studies, The Florida StateUniversity, COAPS Rep. 96-1, 49 pp.

——, M. A. Bourassa, and R. J. Sharp, 1999: Establish-ing more truth in true winds. J. Atmos. OceanicTechnol., 16, 939–952.

——, D. M. Legler, and K. V. Verzone, 2001: Quantify-ing uncertainties in NCEP reanalyses using high-quality research vessel observations. J. Climate, 14,4062–4072.

Smith, W., and Coauthors, 1996: Observations of theinfrared radiative properties of the ocean—Implica-tions for the measurement of sea surface temperaturevia satellite remote sensing. Bull. Amer. Meteor. Soc.,77, 41–51

Sui, C. H., and K. M. Lau, 1997: Mechanisms of short-term sea surface temperature regulation: Observa-tion during TOGA COARE. J. Climate, 10, 465–472.

Verschell, M. A., M. A. Bourassa, D. E. Weissman, andJ. J. O’Brien, 1999: Model validation of the NASAscatterometer winds. J. Geophys. Res., 104, 11 359–11 374.

Weare, B. C., 1989: Relationships between net radiationat the surface and the top of the atmosphere derivedfrom a general circulation model. J. Climate, 2, 193–197.

Webster, P. J., C. A. Clayson, and J. A. Curry, 1996:Clouds, radiation, and the diurnal cycle of sea sur-face temperature in the tropical western Pacific. J.Climate, 9, 1712–1730.

——, and Coauthors, 1998: Monsoons: Processes, pre-dictability and the prospects for prediction. J.Geophys. Res., 103, 14 451–14 510.

Weill, A., and Coauthors, 2003: Toward a new determi-nation of turbulent air–sea fluxes from several me-soscale experiments. J. Climate, 16, 600–618.

Weissman, D. E., and H. C. Graber, 1999: Satellitescatterometer studies of ocean surface stress. J.Geophys. Res., 104, 11 329–11 335.

Weller, R. A., and S. P. Anderson, 1996: Surface meteo-rology and air–sea fluxes in the western equatorialPacific warm pool during the TOGA CoupledOcean–Atmosphere Response Experiment. J. Cli-mate, 9, 1959–1990.

Wentz, F. J., 1997: A well-calibrated ocean algorithm forSSM/I. J. Geophys. Res., 102, 8703–8718.

——, C. Gentemann, D. Smith, and D. Chelton, 2000:Satellite measurements of sea surface temperaturethrough clouds. Science, 288, 847–850.

WGASF, 2001: Intercomparison and Validation ofOcean-Atmosphere Energy Flux Fields. WCRP/SCORWorking Group on Air—Sea Fluxes Series, Rep. 112,WMO/TD 1036, 303 pp.

Zeng, X., M. Zhao, and R. E. Dickinson, 1998:Intercomparison of bulk aerodynamic algorithms forthe computation of sea surface fluxes using theTOGA COARE and TAO data. J. Climate, 11, 2628–2644.

——, ——, ——, and Y. He, 1999: A multi-year hourlysea surface skin temperature dataset derived from theTOGA TAO bulk temperature and wind speed overthe Tropical Pacific. J. Geophys. Res., 104, 1525–1536.

Zhang, Y. C., and W. B. Rossow, 1995: Calculation ofsurface and top of atmosphere radiative fluxes fromphysical quantities based on ISCCP data sets. 2. Vali-dation and first results. J. Geophys. Res., 100 (D1),1167–1197.

Zhou, D. K., and Coauthors, 2002: Thermodynamicproduct retrieval methodology and validation forNAST-1. Appl. Opt., 41, 6957–6967.