mapping evapotranspiration using modis .. (1)

Upload: diego-vega

Post on 14-Apr-2018

223 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    1/17

    Mapping evapotranspiration using MODIS and MM5 Four-Dimensional

    Data Assimilation

    Keunchang Jang a, Sinkyu Kang a,, Jeachul Kim a, Chong Bum Lee a, Taehee Kim b, Joon Kim c,Ryuichi Hirata d, Nobuko Saigusa e

    a Department of Environmental Science, Kangwon National University, Chuncheon 200-701, Republic of Koreab Department of Groundwater and Geothermal Resources, Korea Institute of Geoscience and Mineral Resources, Daejeon 305-350, Republic of Koreac Department of Atmospheric Science, Yonsei University, Seoul 120-749, Republic of Koread Agro-Meteorology Division, National Institute for Agro-Environmental Sciences, 3-1-3 Kannondal, Tsukuba, 305-8604, Japane National Institute for Environmental Studies (NIES), 16-2 Onogawa, Tsukuba, 305-8506, Japan

    a b s t r a c ta r t i c l e i n f o

    Article history:

    Received 26 February 2009

    Received in revised form 9 November 2009

    Accepted 16 November 2009

    Keywords:

    Evapotranspiration

    MODIS

    MM5 FDDA

    Evapotranspiration (ET), the sum of evaporation from soil and transpiration from vegetation, is of vital

    importance in the hydrologic cycle and must be taken into consideration in assessments of the water

    resources of any region. The MODerate resolution Imaging Spectroradiometer (MODIS) sensor offers a

    promising opportunity for estimating daily ET with a 1 km spatial resolution, but is hampered by frequent

    cloud contamination or data gaps from other factors. In this study, 1) a stand-alone ET model was applied

    and tested during clear or partial cloudy sky conditions using MODIS-based inputs of land surface and

    atmospheric data and 2) meteorological simulations by using Four-Dimensional Data Assimilation (FDDA)

    system between MODIS and the 5th Generation Meso-scale Meteorological Model (MM5) was used in

    cloudy conditions to facilitate continuous daily ET estimates. The MODIS ET algorithm modified from Mu

    et al. (2007) is based on the PenmanMonteith equation and was applied to predict ET at flux measurement

    sites. This algorithm considers both the effects of surface energy partitioning processes and environmental

    constraints on ET. We devised gap-filling approaches for MODIS aerosol and albedo data that were identified

    as bottlenecks to determine retrieval rates of insolation and ET. MODIS-derived input variables (i.e.,meteorological variables and radiation components) for estimating ET showed a good agreement with flux

    tower observations at each site. The retrieval rate of MODIS ET doubled at four flux measurement sites after

    gap-filling with negligible compensation was undertaken for accuracy. In spite of the high accuracy of

    MODIS-derived input variables, MODIS ET showed meaningful errors at the four flux measurement sites.

    These errors were mainly associated with errors in the estimated canopy conductance. During clear sky

    conditions, MODIS was used to calculate ET, while the MODIS-MM5 FDDA system provided input variables

    for the calculation of ET under cloudy sky conditions. The performance of the MODIS-MM5 FDDA system was

    evaluated by comparing ET based on MODIS, which showed a good agreement with the MODIS ET for various

    land cover types. Our results indicate that MODIS can be applied to monitor the land surface energy budget

    and ET with reasonable accuracy and that MODIS-MM5 FDDA has the potential to provide reasonable input

    data of ET estimation under cloudy conditions.

    2009 Elsevier Inc. All rights reserved.

    1. Introduction

    Evapotranspiration(ET) is a collective term forall of theprocesses by

    which water in the liquid or solid phase at or near the earth's land

    surfaces becomes atmospheric water vapor. ET includes the sum of

    evaporation of liquid water from surface waters, soil, and surface

    vegetation, as well as transpiration from the tissues of plant leaves. Over

    the entire land surface of the globe, approximately 62% of the

    precipitation returns to the atmosphere with ET from the land's surface

    and evaporation from open-water surfaces (Dingman, 1994; Fisher

    et al., 2005). ET is not only an important component of hydrological

    cycles including precipitation, runoff, streamflow, and soil water

    content, but is also a key variable forthe assessment of water resources

    in any region. ET also has important effects on climate dynamics and

    terrestrial ecosystem productivity because it is closely related to energy

    transfer processes (Fisher et al., 2008; Nishida et al., 2003a). Accurate

    monitoring of the temporal and spatial distribution of ET is critical to

    improving our understanding of energy and hydrologic partitioning

    between the land surface and the atmosphere (Boegh et al., 2002;

    Cleugh et al., 2007; Mu et al., 2007; Venturini et al., 2008 ). For these

    Remote Sensing of Environment 114 (2010) 657673

    Corresponding author. Tel.: +82 33 250 8578; fax: +82 33 251 3991.

    E-mail address: [email protected] (S. Kang).

    0034-4257/$ see front matter 2009 Elsevier Inc. All rights reserved.

    doi:10.1016/j.rse.2009.11.010

    Contents lists available at ScienceDirect

    Remote Sensing of Environment

    j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / r s e

    mailto:[email protected]://dx.doi.org/10.1016/j.rse.2009.11.010http://www.sciencedirect.com/science/journal/00344257http://www.sciencedirect.com/science/journal/00344257http://dx.doi.org/10.1016/j.rse.2009.11.010mailto:[email protected]
  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    2/17

    reasons, many hydro- and agricultural forest-meteorologists have

    strived to accurately estimate ET for decades, but it remains difficult

    to estimate ET with a high reliability at a large scale area.

    Many methods for estimating ET have been developed, and accurate

    estimates of ET arebecoming availablethrough the useof ground-based

    observations (i.e., the eddy covariance method). However, ground-

    based observations cover only a small area at a regional or global scale.

    Many studies estimating the spatial and temporal distribution of ET in

    largerareashave required a large numberof observation sites because ofthe heterogeneity of landscapes and the wide variation in energy

    transfer processes (Nishida et al., 2003a; Wang et al., 2006). However,

    this approach is expensiveand labor intensive, so other approaches have

    attempted to use satellite remote sensing to estimate ET at large scales

    (Nishida et al., 2003a; Shin & An, 2007; Wang et al., 2006). Satellite

    remote sensing, especially the MODerate resolution Imaging Spectro-

    radiometer (MODIS) sensor, offers promising techniques for estimating

    ET with temporally and spatially continuous information over land

    surfaces. The MODIS sensor provides a number of biophysical variables

    from land surfaces and from the atmosphere, allowing ET to be

    estimated. Several recent studies have used various techniques in

    applying MODIS products to spatially and temporally continuous

    monitoring of ET (Cleugh et al., 2007; Mu et al., 2007; Nishida et al.,

    2003b; Venturini et al., 2008; Wang et al., 2006).

    Nishidaet al.(2003b) and Wang et al. (2006) proposed methods for

    estimating the evaporative fraction (EF), defined as the ratio of ET to

    available energy, using land surface temperature (LST) and vegetation

    index derived fromMODIS. Cleugh et al. (2007) developed an algorithm

    based on the PenmanMonteith equation (PM), the resistance-surface

    energy balance and the PenmanMonteith algorithm (RS-PM), to

    estimate ET using MODIS products and flux tower measurement data

    in two different ecosystems. Mu et al. (2007) proposed a modified

    algorithm based on RS-PM (Revised RS-PM)usingboth MODIS products

    and meteorological datasets from the Global Modeling and Assimilation

    Office (GMAO). This algorithm considers canopy conductance and

    environmental constraints (i.e., minimum air temperature, Tmin, and

    vapor pressure deficit, VPD) to calculate ET, and added an evaporation

    term to the soil. Jang et al. (2009) estimated radiation components and

    instantaneous ET using MODIS products and the Revised RS-PMalgorithm using a modification for surface conductance. The PM

    equation is complex and data-demanding, so an alternative, the

    PriestleyTaylor equation (PT), has been widely applied over the last

    three decades. Venturini et al. (2008) proposed a new PT algorithm for

    estimating ET using MODIS atmospheric and land products alone.

    However, optical and thermal satellite data such as MODIS contain

    some limitations, such as cloud contamination, missing data caused by

    algorithm problems (Mu et al., 2007; Zhao et al., 2005) and a low

    frequencyof measurements compared to that of ground-basedobserva-

    tions.Some researchers have used data frommeteorologicalor radiative

    transfer models to overcome these problems. For example, Boegh et al.

    (2004) reported a methodology for estimating ET combining AVHRR

    satellitedata anda high-resolutionweather forecast model (HIRLAM) in

    Denmark. The 5th Generation Meso-scaleMeteorological Model (MM5;Grell et al., 1995) is a particularly promising tool for providing the

    meteorological variables required as input for ET estimation because of

    its ability to assimilate other forms of data with satellite data (e.g.,

    MODIS). Satellite data provide the ability to monitor atmospheric infor-

    mation from outer space, and have been incorporated into weather

    predictions using a data assimilation approach. In previous studies, the

    Four-Dimensional Data Assimilation (FDDA) approach was used to

    incorporate MODIS data in MM5 (hereafter, MODIS-MM5FDDA), which

    caused the enhancement of weather predictions (Xavier et al., 2006;

    Yamazaki & Orgaz, 2005). One of the byproducts of the MODIS-MM5

    FDDA is the ability to provide missing data (i.e., meteorological data)from satellite observations during cloudy condition, which thusenables

    ET to be estimates under cloudy conditions or during any temporal gaps

    in the data.

    In this study, continuous monitoring of daily ET under all sky

    conditions was developed by integrating MODIS data and MODIS-MM5

    FDDA. We first tested the reliability of the MODIS-derived biophysical

    variables for calculating ET using Eqs. (1)(3) (see Section 2.2.1). The

    stand-alone MODIS ET application uses MODIS land surface and

    atmospheric products (see Section 2.2.2). Our next objective was then

    to develop continuous estimates of daily ET by using MODIS products

    which are gap-filled and replaced by meteorological MM5 simulations

    during cloudy conditions (see Section 2.2.3). For clear sky conditions,

    the stand-alone MODIS ET wasevaluated at fourflux measurementsites

    with different biome types in East Asia. The MODIS-MM5 FDDA was

    implemented and tested for its improved prediction, which provided

    meteorological data to estimate daily ET for cloudy sky conditions. In

    this study, the revised RS-PM algorithm (Jang et al., 2009; Mu et al.,

    2007) was applied to estimate ET and its sensitivity to input variables

    and major sources of error were analyzed and discussed for further

    improvement of the algorithm.

    2. Materials and methods

    2.1. Study sites

    Four flux tower sites and one river basin were selected for algorithm

    tests and spatial mapping of daily ET, respectively. To evaluate meteo-

    rological variables and radiation components derived from the MODIS

    and to validate ET retrieved from the MODIS stand-alone algorithm, weused the observed input data and the latent heat flux (E) from four

    eddy covariance flux towers in East Asia. The study sites in Korea in-

    cluded a cooltemperate deciduous forestin Gwangneung Experimental

    Forest(GDK; Kim et al., 2006) and a dry crop farmlandin Haenam(HFK;

    Lee et al., 2003). The Japanese study sites were a cool temperate

    deciduous forest in Takayama Forest (TKY; Saigusa et al., 2008) and a

    Japanese larch forest in Tomakomai (TMK; Hirata et al., 2008). The

    location, dominant species, tree age,and maximum LAIof study sites are

    described in Table 1.

    Tower flux measurements at the GDK and HFK have been con-

    ducted since 2000 and 2002, respectively. Air temperature (Ta) and

    actual vapor pressure (ea) were measured at a height of 40 m above

    the ground at the GDK and 20 m above the ground at HFK using a

    CSAT3 sonic anemometer (Campbell Sci., Inc., USA) and a HMP-35Humicap (Vaisala, Finland),respectively. Radiation components at the

    GDK were measured at a height of 40 m, while those at the HFK were

    Table 1

    Description of the study sites.

    Site ID Location

    (N, E)

    Elevation

    (m)

    Temperaturea

    (C)

    Precipitationa

    (mm)

    Dominant species Tree age

    (year)

    LAIb

    (m2 m2)

    Study periodc No. of data days

    (clear sky days)

    References

    GDK 37. 45, 127.90 340 11.5 1332 Quercus sp., Carpinus sp. 80200 6 20042892006365 434 (88) Kim et al. (2006)

    HFK 34.55, 126.57 13 13.3 1306 Seasonally cultivated crops 1 3 20040012006365 356 (141) Lee et al. (2003)

    TKY 36.15, 137.42 1420 6.5 2275 Betula, Quercus sp. 50 4 20020012004365 1096 (67) Saigusa et al. (2008)

    TMK 42.74, 141.52 140 6.2 1043 Larix Kaempferi 45 9.2 20020012003365 730 (54) Hirata et al. (2008)

    a Annual mean temperature and precipitation from the AsiaFlux website ( http://asiaflux.yonsei.kr/).b Maximum LAI.c

    Data period used in this study.

    658 K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

    http://asiaflux.yonsei.kr/http://asiaflux.yonsei.kr/http://asiaflux.yonsei.kr/http://asiaflux.yonsei.kr/
  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    3/17

    measured at a height of 15 m above the ground using a Kipp & Zonen

    CNR-1 radiometer (Kipp & Zonen, Netherlands). E was measured

    using an Li-7500 open-path H2O/CO2 analyzer (Li-Cor, USA; Kim et al.,

    2006; Lee et al., 2007a,b).

    Flux measurements at the TKY and TMK in Japan have been

    established since 1993 and 2000, respectively. For the TKY, Ta and ea(HMP 233, Vasala, Finland), and radiation components (MR-50, EKO,

    Japan) were measured at a height of 25 m, while these variables at the

    TMK were measured at a height of 40 m above the ground. E wasmeasured using a Li-6262 open- and closed-path H2O/CO2 analyzer(Li-Cor, USA) at the same height (Hirata et al., 2008; Saigusa et al.,

    2008). Both sets of measurements were acquired at average intervals

    of 30 min (Hirata et al., 2008; Kim et al., 2006; Lee et al., 2007a,b; Lee

    et al., 2003; Saigusa et al., 2008).

    To extend the spatial scale with the revised RS-PM and to test the

    applicability of continuous monitoring of ET using the MODIS-MM5

    FDDA system, we selected the Geum river basin that is located in the

    mid-western Korea Peninsula. It covers 14,470 km2 andland usetypes

    in the basin include forests (61.5%), cropland (31.6%), grassland

    (1.4%), and other types (3.2%, i.e., urban, water, and wetland).

    2.2. Algorithms

    2.2.1. Evapotranspiration

    The PM algorithm for computing potential ET has had many suc-

    cessful applicationsin hydrologicaland agricultural foreststudies and in

    a variety of hydroclimaticregimes (Chenet al., 2005; Cleughet al., 2007;

    Federer et al., 1996; Nishida et al., 2003b). The revised RS-PM used in

    thisstudy is a modification of the RS-PM algorithm proposed by Cleugh

    et al. (2007). Mu et al. (2007) improved the RS-PM algorithm in four

    ways: (1) canopy conductance (Cc) is controlled by both daily Tmin and

    VPD; (2) the Leaf Area Index (LAI) is expressed as a scalar to expand

    fromstomatal conductance(gc) at theleaf level to a canopyconductance

    (Cc); (3) replacing the Normalized Difference Vegetation Index (NDVI)

    with the EVI (Enhanced Vegetation Index) in calculating the fractions of

    vegetation cover; and (4) using a soil evaporation term to account for

    areas with sparse canopy cover. Latent heat flux (E; Eq. (1)) was

    divided into canopy ET (Eveg; Eq. (2)) and soil evaporation (Esoil;Eq. (3)) to estimate total ET.

    E = Eveg + Esoil 1

    Eveg =Rn + cpesea = ra + 1 + rS = ra

    2

    Esoil =Rn;soil + cpesea = ra

    + rtot = ra

    RH

    100

    esea =1003

    where (Pa K1) is the slope of the curve relating saturated watervapor pressure (es, Pa) to temperature (K); (Pa K

    1) is the psy-

    chrometric constant; Rn (W m2) is available energy; (kg m3) is air

    density, cp (J kg

    1 K

    1) is the specific heat capacity of air; ea (Pa) is theactual water vapor pressure; ra (s m

    1) is the aerodynamic resistance;

    and the surface resistance (rs, s m1) is the effective resistance to

    transpiration from the plant canopy. More details on the resistance

    terms are summarized in Appendix A.

    Stomatal conductance (gc) was calculated using maximum leaf

    conductance (CL) with environmental controls of daily minimum air

    temperature (Tmin) and vapor pressure deficit (VPD, i.e., esea).Complete details on the scheme for computing stomatal conductance

    and values for the environmental control constraints are described in

    the Biome Properties Look-Up Table (BPLUT)in the MODIS17 Gross and

    NetPrimary Production (GPP/NPP) algorithm (Heinsch et al., 2003). For

    each biome type, the BPLUT provides values of maximum leaf

    conductance (CL), defined as the mean potential stomatal conductance

    per unit leaf area. Our preliminary test on MODIS ET resulted in

    considerable underestimation, compared with flux tower ET. In this

    study, therefore, we applied a different dataset of maximum leaf

    conductance(CL), as suggestedby Federer et al.(1996), for a typical land

    cover type (Table 2).

    We also added a new multiplicative term (fs) in calculating canopy

    conductance (Cc) from stomatal conductance (gc) and LAI (Eq. (4)).

    The fs variable is a shelter factor that accounts for the fact that some

    leaves are sheltered from the sun and wind and thus transpire at

    lower rates (Dingman, 1994). Based on other studies, fs decreasesmonotonically from 1 to 0.5 with no canopy (LAI=0) and closed

    canopy (LAI= 3; Allen et al., 1989; Carlson, 1991; Eq. (5)). We

    calculate fs for LAI changes as shown in Eq. (5):

    1

    rs= Cc = fs LAI gc 4

    fs =fsopen +

    LAILAImin fsclosefsopenLAImaxLAImin

    0 LAIb 3

    fsclose LAI 3

    8>: 5

    where, fs_open and fs_close are shelter factors for sparse and dense

    vegetation, which take on values of 1 and0.5, respectively; LAI, LAImin,

    and LAImax are leaf area indices for current, leafless (no canopy), anddense (closed canopy) vegetation status.

    In this study, Eqs. (1)(3) were utilized to estimate instantaneous

    ET at MODIS overpass time by using inputs from MODIS data for clear

    or partial cloudy condition. The equations were also used for daily ET

    estimation on cloudy days by using meteorological input data from

    MODIS-MM5 FDDA simulation.

    2.2.2. MODIS-derived instantaneous and daily evapotranspiration

    The instantaneous ET (IET, W m2) during the satellite overpass

    was estimated by Eqs. (1)(3) with MODIS-derived input variables.

    The procedure of IET estimation was described in Fig. 1 and equations

    associated with the process were summarized in Appendices AC.

    Input meteorological variables for calculation of IET were either

    extracted directly or estimated from MODIS atmospheric and landproducts. Especially, MODIS-derived net radiation, as a key variable of

    ET estimation was explained in Section 2.2.4.

    Many hydrological applications and models require ET at a daily

    level (mm day1). Nishida et al. (2003a) reported evaporation

    fraction (EF) as an index for ET, which was defined as a ratio of

    latent heat flux (ET, W m2) to available energy (generally equal to

    INR, W m2; Shuttleworth et al., 1989). Other previous studies

    showed that diurnal patterns of the EF were nearly constant during

    the daytime (Nishida et al., 2003a;Shuttleworth et al., 1989). Daytime

    ET can, therefore, be estimated using the instantaneous EF concept

    when daytime average net radiation is known. In this study, we

    further assumed that nighttime ET is negligible and, hence, daytime

    ET is nearly equal to daily ET.

    Bisht et al. (2005) proposed a method to expand instantaneous netradiation (Rn or INR) to daily average net radiation (DANR) for clear sky

    Table 2

    Typical values of maximum leafconductance (CL) forthebiome types used in this study

    (Federer et al., 1996).

    Land cover CL (m s1)

    Conifer forest 0.0053

    Broadleaf forest 0.0053

    Savanna/shrub 0.0053

    Grassland 0.0080

    Tundra/nonforest wetland 0.0066

    Desert 0.0050

    Typical crop 0.0011

    659K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    4/17

    conditions based on the sinusoidal model of Lagouarde and Brunet

    (1983). In this study, we employed a sinusoidal model for estimating

    daily ET, which requires time information for satellite overpass (toverpass),

    sunrise (trise), and sunset (tset).Time information for trise and tset can be

    obtained by using a function of latitude and solar declination angle

    (Dingman, 1994). To expand the model to obtain daily ET, DANR is first

    multiplied with daytime length (Dlength, i.e., tsettrise) to derive dailytotal Rn. Then, daily ET was estimated by multiplying the instantaneous

    EF by the daily total Rn. Thecalculation methods forthe sinusoidal model

    and daily ETcanbe found inAppendixA (seealso Bisht et al., 2005 forthe

    sinusoidal model), and each parameter for the sinusoidal model was

    calculated using the methods suggested by Dingman (1994).

    2.2.3. Continuous monitoring of daily ET

    For clear or partial cloudy days, daily ET was estimated with input

    variables derived from MODIS atmospheric and land products as

    described above (Fig. 1). For cloudy days, the MODIS-derived meteo-

    rologicalinput data were not provided andhence, replaced with those

    from MODIS-MM5 FDDA predictions to estimate cloudy-day daily ET

    (Fig. 2). Detailed information on MODIS-MM5 FDDA can be found in

    Section 2.4.

    Fig. 2 shows a schematic diagram for continuous monitoring of ET

    using either stand-alone MODIS (hereafter, MODIS ET) or the

    combined use of both MODIS land products and MODIS-MM5 FDDA

    (hereafter, MODIS-MM5 ET). The operation of this system isdetermined by sky conditions derived from the clear pixel number

    of MODIS07 (see Section 2.3.1). Under clear and partially clear sky

    conditions (clear pixel5), MODIS was used to calculate MODISinstantaneous anddaily ET (Fig. 1). Forthe cloudy skycondition (clear

    pixelb5), MODIS-MM5 ET was calculated using meteorological

    datasets (i.e., air temperature (Ta), humidity (RH), pressure (P), and

    radiation components (Rs, Rl, and Rn)) derived from the MODIS-MM5

    FDDA and MODIS land products (i.e., vegetation indices and albedo).

    In both cases, the Eqs. (1)(3) were utilized with different input

    variables. The overall processing stream for estimating MODIS-MM5

    ET is identical to those for MODIS ET, but meteorological variables

    produced via the MODIS-MM5 FDDA were used for continuous

    estimation of ET under cloudy conditions or for any missing data

    period. Subsequently, continuous monitoring of daily ET was

    implemented by combining both MODIS ET and MODIS-MM5 ET for

    clear or partially clear days and for cloudy days, respectively.

    2.2.4. Radiation components

    Net radiation (Rn, W m2) is a key variable for understanding energy

    budgets, climate monitoring, weather prediction, and agricultural meteo-

    rology, and is defined as the difference between downward and upward

    radiation components at the land's surface. It can be expressed as:

    Rn = Rs Rs + Rl Rl 5

    Bisht et al. (2005) and Ryu et al. (2008) proposed a method for

    estimating Rn derived entirely from MODIS. In this study, we employed

    the methodology for estimating Rn suggested by Ryu et al. (2008). Rs

    (equalto insolation) is a vital component of land surface energybudgets

    and is used as a key meteorological input variable, forcing ET and

    photosynthesis in many ecosystem and hydrological models (Hwang

    et al., 2008; Kang et al., 2002; Landsberg & Waring, 1997; Running &

    Coughlan, 1988). We used the parameterization scheme for Rs

    developed by Bird and Hulstrom (Clear Sky Model; 1981) to consider

    both diffuse and direct insolation. The Clear Sky Model has been

    evaluated in several studies and produced accurate results for

    estimating insolation (Annear & Wells, 2007; Chen et al., 2007; Ryu

    et al., 2008). Recently, the Clear Sky Model was reevaluated and wasdetermined to be the most accurate clear sky insolation model among

    five alternative insolation models incorporated into water quality

    models (Annear & Wells, 2007). Rs can be described as the proportion

    of Rs that is dependent on the albedo () or shortwave reflectancefrom the ecosystem surface. The albedo () data can be derived fromMODIS43 BRDF products. Prata (1996) proposed a scheme for

    estimating Rl for clear sky days, which outperforms other widely

    used formulas. Bisht et al. (2005) and Ryu et al. (2008) employed this

    scheme in their studies, and reported a high accuracy with ground

    observations. Rl can be calculated using Prata's (1996) scheme based

    on the SteffanBolzmann equation, and we employed Liang's (2004)

    scheme to estimate the surface emissivity using MODIS11 emissivity

    data. More detailed methods for calculating radiation components,

    albedo, and emissivity are described in Appendix B.

    Fig. 1. Processing flow for estimating MODIS-based radiation components and evapotranspiration.

    660 K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    5/17

    2.3. MODIS data processing

    MODIS sensors were launched on board the National Aerodynamics

    and Space Administration (NASA) Earth Observing System (EOS) Terra

    and Aqua satellites on December 1999 and May 2002, respectively. The

    wide spectral range (mainly 36 bands), high spatial and temporal reso-

    lution(250,500, and1000 m andmore thantwiceper day, respectively)

    of MODIS enables MODIS to observe the earth's atmosphere and land

    surface, and to monitor their changes continuously (Bisht et al., 2005;Masuoka et al., 1998; Ryu et al., 2008; Seemann et al., 2003). In this

    study, we used the Aqua-MODIS products to estimate ET ( Fig. 1).

    2.3.1. Atmospheric products

    The MODIS07 atmospheric profile products have a spatial resolution

    of 55 km2 and provide several instantaneous variables, of which

    latitude, longitude, air and dew-point temperature, total ozone burden,

    surface pressure, total precipitable water vapor, and solar zenith angle

    (Seemann et al. 2006) were used in this study. Air and dew-point

    temperatures were extracted from the lowest valid layer among 20

    vertical atmospheric pressure levels and were used to calculate satu-

    rated vapor pressure (es) and actual vapor pressure (ea), respectively

    (see Appendix B).

    The quality assurance (QA) field provides information for the cloudmask, which indicates how many of the 1 1 km2 sub-pixels have clear

    sky conditions within a MODIS07 pixel resolution of 5 5 km2. In other

    words, if the number of 11 km2 clear pixels is equal to 25, the sky

    condition is perfectly clear within a 55 km2 pixel resolution. If the

    number of clear pixel is less than 5, the dataset was not produced.

    Considering the sky conditions(i.e.,perfect clear, partial clear, or cloudy

    sky), we tested the fifth byte of the QA Scientific Data Set (SDS) that is

    assigned each 10 byte array and extracted the clear pixel for the sky

    conditions.

    The MODIS04 aerosol products (10 10 km) provide aerosol

    optical depth (AOD) at three microwave lengths (i.e., 0.47, 0.55, and

    0.67 m). Instantaneous AOD measurements were used as inputvariables to estimate Rs but the retrieval rate is lower than for

    MODIS07 variables, which results in a low retrieval rate for Rs and,

    ultimately, forET. Therefore, we increasedthe retrievalratesof MODIS

    Rs by gap-filling of themissing aerosol data used byJang et al. (2009).

    For the missing aerosol data, we simply used the monthly mean

    aerosol value of each pixel from 2004 to 2006.

    2.3.2. Land products

    The MODIS11A1 products produced by land surface generalized

    split-window algorithm contain land surface temperature (LST) and

    emissivity for bands 31 (10.7812.38 m) and 32 (11.7012.27 m)with a spatial resolution of 11 km2, both of which were used as

    input variables for calculating Rl.

    The MODIS13A2 products contain 16-day composite vegetation

    indices such as NDVI and EVI (Huete et al., 2002). The vegetation

    indices were generally reliable using the 16-day composite method,

    but sometimes showed unlikely fluctuations during the early

    spring or during the summer rainy season in East Asia. For this

    reason, we applied the smoothing technique suggested by Kim et al.

    (2008) to reconstruct smoothed seasonal changes in vegetation

    indices.

    The MODIS43B3 products provide information on 16-day albedo at

    various bands. We used the 10th band of the black and white sky

    albedos at 1 km resolution to calculate both downward (Rs) and

    upward shortwave radiation (Rs). However, MODIS43B3 albedocontains some cloud-contaminated or missing data that was identified

    as a majorfactor contributing to thelow retrieval rate of insolation data

    and ultimately those of the ET and MODIS04 aerosol products. There-

    fore, we reconstitutedcontinuous seasonal changes in albedo by using a

    spatial and temporal gap-filling technique which considered the land

    cover type, as suggested by Kang et al. (2005)and Jang et al.(2009). The

    process has two steps (see Fig. 2 in Kang et al., 2005) including spatial

    and temporal interpolation procedures. First, if albedo datafor any pixel

    was unreliable, as identified by quality control (QC) flags or missing

    values, a fill value for that pixel was derived by spatially interpolating

    surrounding cloud-free pixel values having the same land cover type

    within an overlying 55 pixel window (25 km2). Second, if there was

    no cloud-free pixel within the specified window, a temporal interpo-

    lation was performed using the following simple assumption: when the

    Fig. 2. A schematic diagram for continuous estimation of daily ET using MODIS land products and MODIS-MM5 FDDA meteorological data.

    661K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    6/17

    unreliable or missing pixel from a previous 16-day period was cloud-

    free, thecurrent pixelvalue assumes thevalueof thepreviouscloud-free

    pixel.

    2.3.3. Sensitivity analysis and error statistics

    In this study, we tested the ET model sensitivity to key MODIS

    input variables such as Ta, VPD, LAI, and Rn. To evaluate the sensitivity,

    VPD, LAI, and Rn were varied from 20 to +20% with respect to

    MODIS-derived input variables, but Ta was changed from 2 C to+2 C. The percent change from MODIS-derived ET was calculated toevaluate the model sensitivity.

    Mean Error (ME) and Root Mean Square Error (RMSE) were

    computed to evaluate the reliability of temperature, VPD, radiation

    components, and ET. These error statistics were used to confirm the

    bias for model performance and model accuracy in this study. ME and

    RMSE were computed by the following equations:

    ME =1

    n

    n

    i = 1xsxo

    RMSE =1

    n

    n

    i = 1xsxo

    2 1=2

    where n is the number of samples; and xs and xo are the satellite-

    based and ground observation-based values, respectively.

    2.4. MODIS-MM5 FDDA

    2.4.1. Domains of MM5 simulation

    MM5 Version 3 Release 7 (Grell et al., 1995) was developed and

    maintained by Pennsylvania State University (PSU) and the National

    Center for Atmospheric Research (NCAR). In this study, MM5

    simulations were conducted with two nested domains, and had grid

    resolutions of 30 and 10 km, respectively. Fig. 3 shows the model

    domain setup. The first nested domain (D1), with a spatial resolution

    of 30 km, comprises the eastern part of Asia, including China and

    Japan. The second domain (D2) covers the southern Korean Peninsulaand part of the surrounding seas. The model employed 35 vertical

    layers. The lowest computational layer was approximately 23.6 m

    above ground level (AGL) and the top layer was at a height of 50 hPa.

    The initial field and boundary conditions were extracted from

    Regional Data Assimilation and Prediction System (RDAPS) data

    (Lee et al., 2007a,b), which is originally based on the MM5 meso-scale

    model developed by PSU/NCAR. The RDAPS is the operational model

    in the Korea Meteorological Administration (KMA) and was produced

    from the MM5 model simulation with horizontal resolutions of 30 km

    (191 171 grids) and 33 levels of vertical terrain-influenced coordi-

    nates, centered at 38 N, 126 E in the East Asia region. The other

    major components used in this model include mixed-phase micro-

    physics (Reisner et al., 1998), simple shortwave radiation physics

    with cloud interactive radiation (Dudhia, 1989), the RRTM (RapidRadiative Transfer Model) longwave radiation (Mlawer et al., 1997),

    the Grell convective scheme (Grell et al., 1995) and the MRF planetary

    boundary layer (PBL) scheme (Hong & Pan, 1996).

    2.4.2. Implementation of MM5 FDDA

    During the numerical simulations, large errors can accumulate due

    to the uncertainties in the model. To help control the growth of errors,

    MM5 assimilates observations both at the initial time and continuously

    throughout the integration period. This technique, known as the Four-

    Dimensional Data Assimilation (FDDA) or nudging (Stauffer &

    Seaman, 1994), uses an explicit dynamic predictive model combined

    with observations to follow the evolving atmospheric fluid dynamics.

    FDDA is a suitable technique for incorporating observations or satellite

    data into a forecast model such as MM5. The model's prognostic

    equationscombine current andpast datato provide temporal continuity

    and dynamic coupling among the various fields analyzed.

    We performed data assimilation with selected observational and

    MODIS data. The data assimilation for MM5 has two different nudging

    procedures: analysis vs.observational nudging. In this study, we applied

    two nudging process, simultaneously. Observations from ground

    weather stations and a few radiosonde data in the East Asia region

    were used for analysis nudging, while MODIS atmospheric profile ofair

    temperature was utilized for observational nudging, respectively. Gridmaps of the 20 vertical layers of temperature profile of MODIS07 were

    prepared for the observational nudging. Each grid map has information

    of air temperature of each clear sky pixel at the corresponding pressure

    level.

    We attempted numerical experiments to investigate the impact of

    alternative nudging methods (i.e. without and with MODIS data) on

    MM5 FDDA simulations: one adopted only analysis nudging with a few

    radiosonde and meteorological data, but the other utilized the analysis

    nudging together with the observational nudging using the vertical

    profiles of MODIS temperature, respectively. The latter case is MODIS-

    MM5 FDDA as indicated in this study and the former is conceived as

    conventional MM5 FDDA. The numerical experiments were utilized to

    evaluate the performance of the MODIS-MM5 FDDA by comparison

    with the conventional MM5-FDDA simulation. MODIS observations of

    profiler temperatures were adjusted slightly for the 10 km domain. In

    this process, MODIS provided 20 vertical layers of temperature data for

    clear sky days for the MM5 FDDA. The MODIS-MM5 FDDA system

    provides various meteorological dataset, including air temperature,

    humidity, pressure, and radiation components that are key input

    variables for calculating ET under cloudy conditions.

    3. Results

    3.1. Meteorological variables

    The MODIS-derived air temperature showed a good agreement

    with the four flux tower measurement sites for clear sky conditions

    (Table 3). For the GDK site, the RMSE for 20042006 were between

    2.8 C and 3.6 C, and correlation coefficients (r) were from 0.96 to0.97. HFK also showed a good agreement with a RMSE of 2.5 C and

    rvalues of 0.96 for 3 years (20042006). For the TKY, RMSE and rfor

    20022004 were 2.4 C and 0.95, and those for the TMK were 2.8 C

    and 0.94, respectively. The magnitudes of the errors were generally

    similar to or smaller than those reported in previous studies. Prihodko

    and Goward (1997) estimated air temperature using the relationship

    of surface temperature and vegetation index from AVHRR over

    northeastern Kansas. They reported a RMSE of 2.9 C. Houborg and

    Soegaard (2004) showed RMSE of 2.5 C using Terra-MODIS over an

    agricultural area in Denmark, and Bisht et al. (2005) reported 5.0 C of

    RMSE over the Southern Great Plains, USA.

    Actual vapor pressure (ea), estimated using MODIS07 dew-point

    temperatures, showed a reasonable agreement with the field

    measurements at four sites. The RMSE at the GDK, HFK, TKY, andTMK were 234, 286, 354, and 246 Pa, respectively. These results were

    generallysimilar to or smaller than those reported in Ryu et al. (2008)

    for two flux tower sites in Korea (Terra, 400 Pa and Aqua, 320 Pa).

    Vapor pressure deficit (VPD), defined as the difference between

    saturated vapor pressure and actual vapor pressure, also showed a

    good agreement with the field measurements. However, VPD was

    generally underestimated at the GDK, HFK, and TMK sites, but

    overestimated at the TKY. More details for the error analysis of VPD

    are presented in Table 3.

    3.2. Radiation components

    Forfour flux tower sites and Geum river basin, we applied a simple

    gap-filling approach (Jang et al., 2009) for MODIS aerosol and albedo

    662 K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    7/17

    data that were identified as a bottleneck to determine retrieval rates

    ofRs and Rn for clear sky days. Jang et al. (2009) evaluated the gap-

    filled Rs using ground-based measurement data from 22 stations of

    the Korea National Weather Stations (NWS), which resulted in a

    considerable increase in the retrieval rate of MODIS Rs. In this study,

    even after gap-filling for MODIS aerosol and albedo data, the retrieval

    rates of MODIS Rs for the flux tower sites remained at approximately

    20% or so for Aqua-MODIS. By including partially clear sky conditions

    with the gap-filling techniques for aerosol and albedo, the retrieval

    rate increased from 16 to 44% for the study areas.

    All statistics information for the four gap-filled radiation compo-

    nents and net radiation are summarized in Table 3. Rs showed a good

    agreement (a coefficient of determination of over 0.82) at all sites, butwas generally underestimated. ME and RMSE were 29.3 and69.3 W m2 at the GDK, 38.1 and 54.1 W m2 at the HFK, 46.7and 74.4 W m2 at the TKY, and 12.0 and 85.6 W m2 at the TMK,respectively. Rs was slightly underestimated at the TKY and TMK sites

    with a ME (RMSE) of27.3 (34.8) W m2 and10.6 (19.2) W m2,

    respectively. Rl and Rl were underestimated at both TKY and TMK

    sites. TMK showed a better agreement (ME, 25.5 and8.6 W m2)with the field measurements than TKY (37.3 and35.7 W m2).

    Rn showed a reasonable accuracy with the field measurements

    with RMSE (ME) of 90.5 (38.6), 44.1 (+32.1), 61.2 (29.1), and76.4 (16.6) W m2 at the GDK, HFK, TKY, and TMK, respectively.MODIS Rn was generally underestimated at the GDK, TKY, and TMK,

    but overestimated at the HFK. The errors found in this study were

    comparable with those reported in other studies from Terra-MODIS,

    with a RMSE of 74 W m2 (Bisht et al., 2005) and from Terra and

    Aqua-MODIS with a RMSE of 4765 W m2 (Ryu et al., 2008).

    3.3. MODIS instantaneous and daily ET

    In comparison with flux tower measurements, MODIS-derived

    clear sky instantaneous ET resulted in an overestimation at the GDK

    (ME=+41 W m2) but an underestimation at the HFK (7 W m2),TKY (52 W m2), and TMK (44 W m2) sites (a, b, c, and d in

    Fig. 3. Domains used inthe MM5 model. The circles () andtriangles() indicatethe upper-airstations and NationalWeather Stations(NWS), respectively. (D1)and (D2)show the

    30 km and 10 km domains of MM5 for predictive simulations.

    Table 3

    Means (a) and error statistics(b) forMODIS-derived meteorological variablesand radiationcomponentson clear sky daysat fourflux measurement sites. Here, Ta is air temperature;

    ea, vapor pressure; VPD, vapor pressure deficit; Rs and Rs, downward and upward shortwave radiation; Rl and Rl, downward and upward longwave radiation; and Rn, net

    radiation.

    Site Ta (C) ea (Pa) VPD (Pa) Rs (W m2) Rs (W m

    2) Rl (W m2) Rl (W m

    2) Rn (W m2)

    (a) Means offlux tower measurements with MODIS-derived variables in parenthesis

    GDK 11.0

    (10.0)

    664.3

    (686.8)

    946.9

    (711.2)

    681.9

    (652.6)

    (74.4)

    (285.5)

    (551.4)

    534.8

    (488.5)

    HFK 16.0

    (15.6)

    877.8

    (970.5)

    1129.6

    (920.8)

    766.4

    (728.3)

    (118.2)

    (312.9)

    (417.4)

    474.9

    (507.0)

    TKY 12.5

    (11.7)

    910.7

    (687.6)

    633.2

    (821.1)

    782.9

    (736.4)

    111.0

    (83.7)

    319.8

    (282.5)

    417.2

    (391.6)

    580.3

    (550.7)

    TMK 11.5

    (10.0)

    859.6

    (928.0)

    545.8

    (552.8)

    696.7

    (684.7)

    91.8

    (81.2)

    318.0

    (292.5)

    397.8

    (389.2)

    551.7

    (535.1)

    (b) Mean error (ME) with root mean square error (RMSE) in parenthesis

    GDK 1.0(3.5)

    +22.5

    (233.5)

    222.3(269.1)

    29.3(69.3)

    ()

    ()

    ()

    38.6

    (90.5)

    HFK 0.4(2.5)

    +92.3

    (258.8)

    205.8(387.7)

    38.1(54.1)

    ()

    ()

    ()

    +32.1

    (44.1)

    TKY 0.7(2.4)

    223.1(347.9)

    +188.0

    (188.0)

    46.7(74.4)

    27.3(34.8)

    37.3(45.3)

    35.7(30.9)

    29.1(61.2)

    TMK 1.5(2.8)

    148.3(243.3)

    +7.0

    (293.4)

    12.0(85.6)

    10.6(19.2)

    25.5(28.9)

    8.6(14.8)

    16.6(76.4)

    663K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    8/17

    Fig. 4). HFK showed more scattered relationships (r=0.60 and

    RMSE= 84 W m2) than GDK (0.76 and 75 W m2), TKY (0.66 and

    119 W m2), and TMK (0.74 and 91 W m2), respectively (Table 4).

    Uncertainty in MODIS-derived meteorological data (Table 3) affects

    errors in the stand-alone MODIS-derived instantaneous ET. Hence, the

    tower observed meteorology (i.e. air temperature, vapor pressure, and

    net radiation) were utilized together with MODIS vegetation indices to

    test reliability of the ET algorithm suggested in this study. Except for

    GDK site(ME=+64 W m2), the useof tower meteorology resulted in

    better agreements with the flux tower ET at HFK (+2 W m2), TKY

    (10W m2), and TMK (0.3 W m2), respectively (Table 4). Thisindicates that our ET algorithm estimates more reliable ET with more

    accurate meteorological data but still contains other source of

    uncertainty in addition to the input meteorology.

    Fig. 4. Comparison of MODIS-derived instantaneous ET (W m2) and flux tower measurements. Left images show instantaneous ET for clear sky conditions (clear pixel=25) and

    right images show instantaneous ET for the clear and partial clear sky conditions (clear pixel5) at the GDK (a, e), HFK (b, f), TKY (c, g), and TMK (d, h).

    664 K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    9/17

    We also compared MODIS instantaneous ET with flux measure-

    ments for all sky conditions, including both perfect and partial clear

    sky conditions (MODIS07 clear pixel N=5; e, f, g, and h in Fig. 4).

    Including partial clear sky conditions did not significantly reduce the

    accuracies of MODIS ET at the four sites. The ME and RMSE at the GDK

    and TKY and ME at the HFK decreased rather than increased for data

    with clear sky conditions, while the ME and RMSE at the TMK and

    RMSE at the HFK increased (Table 4). By including partial clear sky

    conditions, the retrieval rates of MODIS ET increased from 21 to 35% at

    the GDK,from 16to 27% attheHFK,from 8 to17%at the TKY,andfrom

    10 to 19% at the TMK, respectively, with negligible compensations in

    accuracy.Instantaneous ET (W m2) was converted to daily ET (mm day1)

    using our scaling method in Appendix A.3. Comparisons between

    MODIS daily ET and flux tower measurements showed a better agree-

    ment (Fig. 5) thanthose for theMODIS instantaneous ET(Fig. 4). MODIS

    daily ET resulted in an overall overestimation at the GDK with ME

    of +0.48 mm day1, but an underestimation at the HFK (0.19 mmday1), TKY (0.53 mm day1), and TMK (0.56 mm day1). Aswell, a better linearity was observed for daily ET with the flux tower

    data for the GDK (r=0.90, RMSE=0.79 mm day1), HFK (0.81 and

    Table 4

    Errorstatistics (MEwith RMSEin parenthesis)for Aqua-MODISinstantaneousET (W m2)

    anddailyET (mmday1) atflux sites.InstantaneousET analyzed forclear skyand partially

    clear sky conditions, and daily ET analyzed for clear sky conditions.

    Site Instantaneous ET

    for clear skya

    (W m2)

    Instantaneous ET

    for clear skyb

    (W m2)

    Instantaneous ET for

    partially clear skyb

    (W m2)

    Daily ETb

    (mm day2)

    GD K +63.6 (100.0) +41.0 ( 75.4) +30.8 ( 73.3) +0.06 ( 0.19)

    HFK +2.4 (79.0) 7.1 ( 84.1) +2.8 ( 87.4) 0.19 (0.79)TKY 10.2 (109.4) 52.1 (119.2) 32.1 (99.3) 0.53 (0.88)TMK 0.3 (77.4) 44.7 (91.6) 59.1 (132.6) 0.56 (1.01)a ET estimated using flux tower meteorological data and MODIS vegetation indices

    product with the revised RS-PM algorithm.b ET estimated using MODIS atmospheric and land products with the revised RS-PM

    algorithm.

    Fig. 5. Comparison of MODIS dailyET (mm day1) estimatedusing a sinusoidalmodel andflux towermeasurements forthe clear skycondition atthe GDK(a),HFK (b), TKY(c), and

    TMK (d).

    665K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    10/17

    0.79 mm day1), TKY (0.84 and 0.88 mm day1), and TMK (0.81 and

    1.01 mm day1) than those of instantaneous ET.

    3.4. MODIS-MM5 ET

    Geographical dependence on model predictions is instructive, and

    thus was investigated through calculating and mapping the main

    statistics. Fig. 6 shows the spatial distribution of correlation

    coeffi

    cients (r) for air temperature (Temp, equal to Ta) and specifi

    chumidity (Q2) at a 2 m height and wind speed (WS) at a 10 m height

    for numerical experiments before and after MODIS-MM5 FDDA.

    Correlation coefficients (r) after MODIS-MM5 FDDA (a, c, and e in

    Fig. 6) were higher than those from the prediction of MM5 alone (b, d,

    and f in Fig. 6). These results indicated that the FDDA system can

    simulate the temporal and spatial variability for near-surface with

    improved accuracy.

    The improved meteorological information produced by MODIS-

    MM5 FDDA was combined with MODIS land product information to

    calculate MODIS-MM5 daily ET (Fig. 3). The MODIS-MM5 ET was used

    for cloudy days to provide a scheme for continuous monitoring of

    daily ET for all sky conditions. First of all, we compared MODIS-based

    daily ET (i.e., MODIS ET) with FDDA-based daily ET (i.e., MODIS-MM5

    ET) to evaluate the predictive abilities of these two techniques. Fig. 7

    shows time series and scatter plots between MODIS ET and MODIS-

    MM5 ET for different land cover types (i.e., forest, cropland, and

    grassland) over the Geum river basin on dates when the MODIS ET

    was retrieved. The time series pattern of MODIS-MM5 daily ET

    showed a good agreement with the pattern of MODIS daily ET for all

    land cover types. This suggests that FDDA-based products can be

    merged with MODIS-based daily ET and can be used to make up for

    weak points due to missing MODIS data during cloudy days. We used

    this technique, and produced accumulated ET over the Geum river

    basin in 2006 (Fig. 8). The average of annual accumulated ET over the

    Geum river basin was approximately 750 mm y1, and showed major

    spatial variations among land cover types. The accumulated ET

    (average 1000 mm) for cropland was generally higher than other

    land cover types, including forests and grasslands. These trends could

    be due to the effect for the relatively high value of maximum leafconductance (CL=11.0 mm s

    1) for croplands.

    3.5. Sensitivity analysis

    We also conducted a model sensitivity analysis on atmospheric

    and land surface data derived from MODIS. Fig. 9 shows the model

    sensitivity (relativepercent change) to the key input variables (i.e., Ta,

    VPD, LAI, and Rn) on the estimation of ET using the Revised RS-PM

    algorithm. A relative percent change in ET of up to 12% was caused

    by changing Rn with 20% at each site. Changes in ET of 911% were

    found for 20% changes in LAI. Changes of Rn and LAI showed a linear

    relationship with ET percent changes as expected from the algorithm.

    VPD and Ta, however, showed nonlinear relationships with ET. The

    sensitivity of ET to VPD was relatively low because of positive effectsof VPD on forced evaporation and negative effects on stomatal

    opening. Our results based on the Revised RS-PM algorithm indicate

    that for the deciduous sites (the GDK and TKY), the positive effect was

    more than offset by higher negative impact on stomatal opening. Tawas associated with many terms of the ET process, including Rl, ,VPD, and stomatal and aerodynamic conductance. For the two

    deciduous sites, the positive effects of Ta on VPD and stomatal

    conductance were offset by adverse effects on net radiation, which

    resulted in a lower sensitivity than at the HFK and TMK sites.

    4. Discussion

    In this study, we have developed a stand-alone algorithm to

    estimate instantaneous and daily ET using only MODIS products. The

    stand-alone algorithm is based on the Revised RS-PM proposed by Mu

    et al. (2007) but does not require any input variables other than

    MODIS products and estimated ET with reasonable accuracy. Also, the

    parameterization of maximum leaf conductance (CL) and the addition

    of the multiplicative term (fs) in the calculation for canopy con-

    ductance (Eq. (4)) implicitly accounts for the sheltering effect caused

    by phenological change, and reproduced the improvement of MODIS

    instantaneous ET. Various meteorological variables (i.e., Ta, ea, and

    VPD) derived by MODIS products showed a good correlation withground-based observations (Table 3) for clear sky days. Improve-

    ments in MODIS products for estimating ET increased the retrieval

    rate (%). In particular, enhancement of MODIS04 aerosol and

    MODIS43 albedo products using the gap-filling method led to

    increases in the retrieval rate for Rs, Rn, and ET and resulted in a

    good agreement with the flux measurements.

    The framework presented here that is used to estimate ET is

    applicable for clear sky conditions only. For continuous monitoring of

    land surface radiation components and ET, however, the average

    retrieval rate of 20% for clear sky conditions still requires further

    improvements. MODIS-derived ET for partially clear sky conditions

    showed similar errors to those for clear sky conditions, while retrieval

    rates nearly doubled (Fig. 4).

    In this study, the MODIS-MM5 FDDA was implemented to provide

    meteorological data for cloudy conditions for continuous monitoring

    using satellite data. Other recent efforts for satellite monitoring of

    land surface radiation components and ET during all sky conditions

    are worthy of note. Tang et al. (2006) suggested a direct method for

    estimating shortwaveradiation (Rs) using multispectral narrowband

    data of MODIS for all sky conditions. They reported that RMSE to Rs is

    less than 20 W m2 for clear sky conditions and 35 W m2 for cloudy

    conditions. Liang et al. (2006) reported the look-up table approach to

    estimate incident photosynthetically active radiation (PAR) using

    MODIS. These approaches were applied using top-of-atmosphere

    (TOA) radiance, which contains information on both the atmosphere

    and surface. Comparison of their PAR results with ground data at

    seven FLUXNET sites yielded the average relative error (%), which

    ranges from 4.1 to 21.9%. Also, Jutla and Islam (2007) presented a

    method for estimating MODIS-based ET using the PriestleyTaylorequation under all sky conditions, and Jutla and Islam (2008)

    suggested that MODIS06 cloud fraction and cloud optical depth

    products could be applied to estimate instantaneous shortwave

    radiation (Rs) for all sky conditions.

    In spite of the high accuracy of the input variables for estimating ET

    (Table 3), the comparison with flux measurement ET showed

    meaningful errors in MODIS-derived ET (Table 4). The causes of these

    errors might be related to operational limitationsof MODIS (i.e., a lower

    temporal resolution, such as 16-day MODIS products or a coarse spatial

    resolution over heterogeneous land cover and/or complex terrain), as

    well as the algorithm or algorithm parameters for estimating ET. Mu

    et al. (2007) discussed the limitations of the parameters in their

    algorithm, including: (1) effects caused by the aerodynamic resistance

    from plant transpiration and the aerodynamic resistance from soilevaporation and (2) the refinement of surface resistance. In this study,

    we tested the effect of the canopy conductance against flux measured

    canopy conductance, which is equal to the inverse of surface resistance.

    Canopy conductance is one of main parameters used to estimate ET

    and is controlled by VI and environmental constraints (i.e., Tmin and

    VPD) in this study. Fig. 10 shows the comparison of differences in

    instantaneous ET and canopy conductance in the MODIS and flux

    measurements at the GDK, HFK, TKY, and TMK. The coefficients of

    determination (r2) were 0.74 at the GDK, 0.71 at the HFK, 0.89 at the

    TKY, and 0.66 at the TMK. This result suggests that canopy conductance

    significantly affected the error of the instantaneous ET. In this study,

    rather than using the MODIS15 leaf area index (LAI) product, we used

    LAI derivedfrom the 16-day MODIS13vegetation index product (NDVI)

    proposed by Fisheret al. (2008) to obtain canopy conductance.We used

    666 K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    11/17

    Fig. 6. Each map shows distributions of correlation coefficients (r) between weather stations and MM5. The size of each circle () is scaled from the minimum (0.15) to the

    maximum values (1) for each map. (a), (c), and (e) showed the result of MM5 only, and others showed the result of MODIS-MM5 FDDA.

    667K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    12/17

    Fig. 7. Comparisons of MODIS-based and FDDA-based daily ET for land cover types over the Geum river basin. The rectangles () and dashed lines () in left images indicate

    MODIS-based and FDDA-based daily ET, respectively.

    Fig. 8. Annual accumulated ET (mm)map derived usingthe mergence technique betweenMODIS and FDDA datasets under cloudy sky conditions over the Geumriver basinin 2006.

    668 K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    13/17

    Fig. 9. The model sensitivity (relative percent change, %) to the change rate of the key input variables on the estimation of ET at the GDK (a), HFK (b), TKY (c), and TMK (d). Ta, VPD,

    LAI, and Rn indicate air temperature, vapor pressure deficit, lead area index, and net radiation, respectively.

    Fig. 10. Comparison of the difference between measured and MODIS instantaneous ET (W m2) and the difference between measured and MODIS canopy conductance (CC) at the

    GDK (a), HFK (b), TKY (c), and TMK (d).

    669K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    14/17

    this product becausethe MODIS15 LAI product produced by the MODIS

    Standard algorithm tends to be higher than ground observations

    (Cleugh et al., 2007; Heinsch et al., 2006; Mu et al., 2007; Wang et al.,

    2004). Canopy conductance is influenced by environmental constraints

    (i.e., Tmin and VPD) and phenological changes. For the relatively long

    (i.e., 16-day) periodof this study, phenologicalchanges at thebeginning

    and end ofthe seasonsmaybe difficultto detectfor biome type.Evenso,

    we found that there was a bias, as the difference between canopy

    conductance valuescannot be explainedby thebias in instantaneousET.Hence, this analysis indicates that a more extensive parameterization of

    canopy conductance and evaluation of aerodynamic resistance for the

    vegetation and soil components are required to improve the prediction

    ability of the revised RM-PM algorithm.

    The MODIS07 air temperature profiles (50 to 1000 hPa) were ex-

    tractedas grid cells forthe FDDA. These cells were then interpolatedinto

    35 sigma pressure levelsof MM5. Thedata assimilation betweenMODIS

    and MM5 showed improvement for predicting simulations of MM5

    (Fig. 6). In particular, air temperature and humidity data used in the

    Revised RS-PM showed an enhancement for spatial and temporal

    distributions. Theimproved data setvia theMODIS-MM5 FDDA used for

    estimating ET at scales greater than the basin scale and the FDDA scale

    was suggested as a solution for obtaining data under cloudy conditions.

    We have produced the daily ET map, which ranges from 480 to

    1200 mm, by exploiting MODIS land products and meteorological

    variables (i.e., air temperature (Ta), humidity (RH), pressure (P), and

    radiation components (Rs, Rl, and Rn)) derived from the MODIS-MM5

    FDDA over the Geum river basin (Fig. 8). The use of combined MODIS-

    based and FDDA-based daily ET for cloudy days or missing pixels

    showed the potential to compensate for weaknesses in optical and

    thermal satellite data such as MODIS. The accordance between

    MODIS-based ET and FDDA-based ET showed a high reliability

    (Fig. 7), suggesting that the FDDA-based ET might show uncertainties

    similar to those of MODIS-based ET in comparison with flux ET

    measurements. The procedure of FDDA-based ET also followed the

    same revised RS-PM algorithm with MODIS land products. Thus, any

    improvement on ET estimations at either the site or the regional scale

    should address the enhancement of ET algorithms, more accurate

    input variables, and robust parameterization (especially for aerody-namic and canopy conductance).

    5. Conclusion

    In this study, we used various Aqua-MODIS products, including

    atmospheric (air and dew-point temperature, total ozone burden,

    surface pressure, total precipitable water vapor, and aerosol optical

    depth) and land surface data (LST, emissivity, VI, and albedo) to

    estimate ET. A major weakness of passive satellite data was the

    relatively low retrieval rate due to cloud cover. In particular, aerosol

    and albedo products were identified as bottlenecks to ultimately

    determine retrieval rates of Rs and ET. Devising simple gap-filling

    approaches for aerosol and albedo products improved (nearly

    doubled) the retrieval rate of Rs and ET (Table 3).The input variables of ET retrieved from MODIS showed a good

    agreement with ground observations for clear sky conditions. The

    comparison with flux measurement ET showed meaningful errors in

    MODIS-derived ET at the four flux measurement sites. The cause of

    this bias was likely the difference in canopy conductance (Fig. 9). A

    more enhanced parameterization of conductance data in algorithms is

    required.

    The approach for continuous monitoring of ET using MODIS land

    products and MODIS-MM5 FDDA was successfully implemented for

    the Geum river basin using satellite data for 2006. For the cropland,

    annual accumulated ET (mm) showed a higher value compared to

    other vegetation cover types. The high ET was influenced by the

    maximum leaf conductance for the cropland. Our current algorithm,

    however, is more suitable for dry crops than for rice paddy fields,

    which is a dominant crop in East Asia. Most of all, consideration of

    evaporation from the free water surface in rice paddy soils caused by

    irrigation is be needed to improve the prediction.

    Our results indicate that MODIS can be applied to monitor land

    surface energy budgets and ET with reasonable accuracy and the

    MODIS-MM5 FDDA has the potential to provide reasonable input data

    for ET estimation under cloudy conditions. However, the issues of

    spatial scale mismatch (resolution) between flux footprints, MODIS,

    and MM5, as well as the temporal scale mismatch (time lag) stillremain and should be the focus of future research.

    Acknowledgements

    We greatly appreciated the comments and editorial correctionsfrom

    Jeffrey Owen. This research was supported by grants from the

    Sustainable Water Resources Research Center of the 21st Century

    Frontier Research Program (Grant Code: 1-8-3), the Basic Research

    Project(07-3211) of KIGAM,the InnovativeForest DisastersR&D Center

    (S210809L010140) of Korea Forest Service, and the A3 Foresight

    Program (CarboEastAsia) of KOSEF.

    Appendix A. Calculation of ET parameters and daily ET

    A.1. Canopy resistance

    1

    rs= Cc = fs LAI gc A1

    gc = CL mTmin mVPD A2

    mTmin =

    1:0 Tmin Tmin openTminTmin close

    Tmin openTmin closeTmin close b Tmin b Tmin close

    0:1 Tmin Tmin close

    8>>>>>>>:

    A3

    mVPD =

    1:0 VPD VPDopenVPDcloseVPD

    VPDcloseVPDopenVPDopen b VPD b VPDclose

    0:1 VPD VPDclose

    8>>>>>>>:

    A4

    A.2. Aerodynamic resistance

    ra =rc rrrc + r

    A5

    rc = rtotc rcorr A6

    rtotc= 107:0 A7

    rcorr =1:0

    273:15 + Ta293:15

    1:75

    101300P1

    A8

    rr = cp

    4:0

    T3

    a

    A9

    670 K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    15/17

    A.3. Calculation of daily ET (DET) using a sinusoidal model

    Rnt = Rn;max sinttrise

    tsettrise

    A10

    Rn;max =INR

    sintoverpasstrise

    tsettrise

    h i A11

    DANR =t sett riseRntdtt sett risedt

    =2Rn;max

    =

    2INR

    sintoverpasstrise

    tsettrise

    h i A12

    DET = DANR Dlength IET

    INR

    0:0036

    2:5A13

    where 0.0036 and 2.5 are unit conversion coefficients.

    Appendix B. Calculation of radiation components

    B.1. Shortwave radiation (bird's clear sky model)

    B.1.1. Direct irradiance

    Idir = I0cosz0:9662TRTOTUMTWTAA14

    TR = expf0:0903M

    0:841 + M

    M1:01g A15

    TO = 10:1611XO1 + 139:48XO0:3035

    0:002715XO1 + 0:044XO + 0 :0003X2O1

    A16

    XO = UOM A17

    TUM = exp0:0127M

    0:26 A18

    TW = 12:4959XW1 + 79:034XW0:6828

    + 6 :385XW1

    A19

    XW = UWM A20

    TA = exp0:873A 1 + A

    0:7088A M

    0:9108A21

    A = 0:2758A;0:38 + 0:35A;0:50 A22

    M = cosz+ 0:1593:885z1:251 A23

    M

    = MP= P0 A24

    B.1.2. Diffuse irradiance

    Idif = I0cosz0:79TOTWTUMTAA0:51TR + Ba1TAS1M + M1:02

    A25

    TAA = 1K11M + M1:06

    1TA A26

    TAS = TA = TAA A27

    Ba = 0:51 + cosz A28

    B.1.3. Total solar irradiance

    Rs = Idir + Idif = 1gsA29

    S = 0 :0685 + 1Ba1:0Tas A30

    Tas = 100:045M0:7

    A31

    B.2. Longwave radiation

    B.2.1. Downward longwave radiation

    Rl = aT4

    a

    A32

    a = 11 + exp 1:2 + 30:5

    n o

    A33

    = 46:5eaTa

    A34

    ea = esTd = 0 :611 exp17:3 Td

    Td + 237:3

    A35

    B.2.2. Upward longwave radiation

    Rl = sT4

    s

    A36

    s = 0:273 + 1:778311:80731321:03732 + 1:774232 A37

    Appendix C. Notations in Appendices A and B

    g ground albedo

    s sky albedoBa ratio of the forward-scattered to the total scattered

    irradiance due to aerosols (0.84)

    Cc canopy conductance (mm s1)

    CL maximum leaf conductance (mm s1)

    cp specific heat capacity of air (J kg1 K1)

    DANR daily average net radiation (MJ day1)

    DET daily evapotranspiration (mm day1)

    Dlength daytime length in hour (i.e. trisetset)ea actual vapor pressure (Pa)

    a air emissivity

    31 emissivity in MODIS band 3132 emissivity in MODIS band 32

    es saturated vapor pressure (Pa)

    s surface emissivity

    fs shelter factor

    gc stomatal conductance (mm s1)

    Idif diffuse irradiance (W m2)

    Idir direct irradiance (W m2)

    IET instantaneous evapotranspiration (W m2)

    INR instantaneous net radiation (W m2)

    I0 solar constant (1353.0 W m2)

    K1 constant of aerosol absorptance (0.1)

    LAI leaf area index (m2 m2)

    air density (kg m3)

    M air mass (g cm1

    )M pressure-corrected air mass

    P surface pressure (mb)

    P1 atmospheric pressure (Pa)

    P0 normal atmosphere (1013 mb)

    ra aerodynamic resistance (s m1)

    rc resistance to convective heat transfer (s m1)

    rcorr correction coefficient for rtotc

    Rl downward longwave radiation (W m2)

    Rl upward longwave radiation (W m2)

    Rn net radiation (W m2)

    Rn,max daily maximum net radiation (W m2)

    rr resistance to radiative heat transfer (s m1)

    rtot total aerodynamic resistance (s m1)

    rtotc constant for rtot(107 s m1

    )

    671K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    16/17

    RS Total solar irradiance (W m2)

    SteffanBoltzmann const, 5.67108 (W m2 K4)

    Ta air temperature (C)

    TA aerosol transmittance

    TAA transmittance of aerosol absorptance

    TAS transmittance of aerosol scattering

    Tas dry air transmittance

    A broadband aerosol optical depth in a vertical path

    A,0.38 AOD in a vertical path at 0.38 mA,0.50 AOD in a vertical path at 0.50 mTd dew-point temperature (C)

    Tmin minimum temperature (C)

    TO transmittance of ozone absorption

    toverpass MODIS overpass time

    TR transmittance of Rayleigh scattering

    trise local sunrise time (h)

    Ts surface temperature (C)

    tset local sunset time (h)

    TUM transmittance of uniformly mixed gases

    TW transmittance of water vapor absorption

    UO ozone amount (atm cm)

    UW precipitable water in a vertical path (cm)

    VPD vapor pressure deficit (Pa)

    precipitable water content of atmosphereXO total amount of ozone in slanted path (cm)

    XW precipitable water in a slanted path (cm)

    z solar zenith angle (degree)

    ReferencesAllen, R., Jensen, M. E., Wright, J. L., & Burman, R. D. (1989). Operational estimates of

    reference evapotranspiration. Agronomy Journal, 81, 650662.Annear, R.L., & Wells,S.A. (2007). A comparisonoffivemodelsfor estimating clear-skysolar

    radiation. Water Researches Research, 43, W10415. doi:10.1029/2006WR0050055.Bird, R. E., & Hulstrom, R. L. (1981). A simplified clear sky model for direct and diffuse

    insolation on horizontal surfaces. SERI Report TR-642-761: Golden, USA.Bisht, G., Venturini, V., Islam, S., & Jiang, L. (2005). Estimation of the net radiation using

    MODIS (Moderate Resolution Imaging Spectroradiometer) data for clear sky days.Remote Sensing of Environment, 97, 5267.

    Boegh, E., Soegaard, H., & Thomsen, A. (2002). Evaluating evapotranspiration rates andsurface conditions using Landsat TM to estimate atmospheric resistance and surfaceresistance. Remote Sensing of Environment, 79, 329343.

    Boegh, E., Soegaard, H., Christensen, J. H., Hasager, C. B., Jensen, N. O., Nielsen, N. W., et al.(2004).Combiningweather prediction andremote sensingdata forthe calculationofevapotranspiration rates: Application to Denmark. International Journal of RemoteSensing, 25(13), 25532574.

    Carlson, T. (1991).Modeling stomatal resistance:An overviewof the1989 workshop at thePennsylvania State University. Agricultural and Forest Meteorology, 54, 103106.

    Chen, J., Chen, X., Ju, W., & Geng, X. (2005). Distributed hydrological model for mappingevapotranspiration using remote sensing inputs. Journal of Hydrology, 305, 1539.

    Chen, R., Kang, E., Ji, X., Yang, J., & Wang, J. (2007). An hourlysolar radiation model underactual weather and terrain conditions: A case study in Heihe river basin. Energy, 32,11481157.

    Cleugh, H. A., Leuning, R., Mu, Q., & Running, S. W. (2007). Regional evaporation estimatesfromflux tower andMODISsatellitedata. RemoteSensing of Environment, 106, 285304.

    Dingman,S. L. (1994).Physicalhydrology, 2ndEdition.New Jersey:PrenticeHall pp.272324.Dudhia, J. (1989). Numerical study of convection observed during the winter monsoon

    experiment using a meso-scaletwo-dimensionalmodel.Journal of AtmosphericSciences,

    46, 30773107.Federer, C.A., Vorosmarty,C., & Fekete,B. (1996). Intercomparisonof methodsfor calculatingpotential evaporation in regional and global water balance models. Water ResourcesResearch,32(7), 23152321.

    Fisher, J. B., DeBiase, T., Qi, Y., Xu, M., & Goldstein, A. H. (2005). Evapotranspirationmodels compared on a Sierra Nevada forest ecosystem. Environmental Modeling &Software, 20, 783796.

    Fisher, J. B., Tu, K. P., & Baldocchi,D. D. (2008). Global estimation of the landatmospherewaterflux basedon monthlyAVHRR andISLSCP-IIdata,validatedat 16 FLUXNET sites.Remote Sensing of Environment, 112, 901919.

    Grell,G. A.,Dudhia,J., & Stauffer,D. R. (1995). A description ofthe fifth-generation PennState/NCAR Mesoscale Model (MM5). NCAR Tech. Note NCAR/TN-3981STR, 117.

    Heinsch, F. A., Reeves, M., Votava, P., Kang, S., Milesi, C., & Zhao, M. (2003). User's guide,GPP and NPP (MOD17A2/A3) products, NASA MODIS Land Algorithm (pp. 157).Missoula, MT: The University of Montana.

    Heinsch, F. A., Zhao, M., Running, S. W., Kimball, J., Nemain, R., Davis, K., et al. (2006).Evaluation of remote sensing based terrestrial productivity from MODIS usingregionaltower eddyflux network observations. IEEE Transactions on Geoscience andRemote Sensing, 44(7), 19081925.

    Hirata, R., Saigusa, N., Yamamoto, S., Ohtani, Y., Ide, R., Asanuma, J., et al. (2008). Spatialdistribution of carbon balance in forest ecosystems across East Asia. Agricultural andForest Meteorology, 148(5), 761775.

    Hong, S. -Y.,& Pan, H. -L. (1996). Nonlocal boundary layer vertical diffusionin a medium-range forecast model. Monthly Weather Review of AMS, 124, 23222339.

    Houborg, R. M., & Soegaard, H. (2004). Regional simulation of ecosystem CO2 and watervapor exchange for agricultural land using NOAA AVHRR and Terra MODIS satellitedata. Application to Zealand, Denmark. Remote Sensing of Environment, 93, 150167.

    Huete, A., Didan, K., Miura, T., Rodriguez, E., Gao, X., & Ferreira, L. (2002). Overview of theradiometric and biophysical performance of the MODIS vegetation indices. RemoteSensing of Environment, 83, 195213.

    Hwang, T., Kang, S., Kim, J., Kim, Y., Lee, D., & Band, L. (2008). Evaluating drought effecton MODIS Gross Primary Production (GPP) with an eco-hydrological model in themountainous forest, East Asia. Global Change Biology, 14, 10371056.

    Jang, K., Kang, S., Kim, H., & Kwon, H. (2009). Evaluation of shortwave irradiance andevapotranspiration derived from Moderate Resolution Imaging Spectroradiometer(MODIS). AsiaPacific Journal of Atmospheric Sciences, 45(2), 233246.

    Jutla, A., & Islam, S. (2007). An algorithm for estimation of evapotranspiration for all skyconditions. EOS Transactions, 88(23) AGUH32A-08.

    Jutla, A., & Islam, S. (2008). Estimation of instantaneous shortwave radiation for all skyconditions using MODIS. Geophysical Research Abstracts, 10 EGU2008-A-10377.

    Kang,S., Kim, S.,& Lee, D. (2002).Spatial andtemporalpatternsof solar radiationbasedontopographyand air temperature. Canadian Journalof Forest Research,32(3), 487497.

    Kang, S.,Running, S. W.,Zhao, M.,Kimball, J.S., & Glassy, J. (2005). Improvingcontinuityof MODIS terrestrial photosynthesis products using an interpolation scheme forcloudy pixels. International Journal of Remote Sensing, 26(8), 16591676.

    Kim, J., et al. (2006). HydroKorea and CarboKorea: Cross-scalestudies of ecohydrology andbiogeochemistry in a heterogeneousand complex forest catchmentof Korea. EcologicalResearch, 21, 881889.

    Kim, S., Kang, S., & Lim, J. (2008). Onset date of forest canopy detected from MODIS leaf

    area index. Journal of Ecology and Filed Biology, 31(2), 153159.Lagouarde, J. P., & Brunet, Y. (1983). A simple model for estimating the daily upward

    longwave surface radiation flux from NOAA-AVHRR data. International Journal ofRemote Sensing, 14(5), 907925.

    Landsberg, J. J., & Waring, R. H. (1997). A generalised model of forest productivity usingsimplified concepts of radiation-useefficiency, carbonbalanceand partitioning. ForestEcology and Management, 95, 209228.

    Lee, H. C.,Hong, J.,Cho,C. -H., Choi, B. C.,Oh, S. -N., & Kim, J. (2003). Surface exchangeofenergy and carbon dioxide between the atmosphere and a farmland in Haenam,Korea. Korean Journal of Agricultural and Forest Meteorology, 5, 6169.

    Lee, D.,Kim,J.,Kim,S. -J., Moon,S.-K.,Lee,J., Lim, J.-H.,et al.(2007).Lessonsfromcross-scalestudies of water and carbon cycles in the Gwangneung forest catchmentin a complexlandscape of monsoon Korea. Korean Journal of Agricultural and Forest Meteorology , 9,149160.

    Lee, S. -H., Kim, Y. -K., Kim, H. -S., & Lee, H. -W. (2007). Influence of dense surfacemeteorological data assimilation on the prediction accuracy of ozone pollution in thesoutheastern coastal area of the Korean Peninsula. Atmospheric Environment, 41(21),44514465.

    Liang,S. (2004).Quantitativeremote sensingof land surface. NewJersey:JohnWiley& Sons.Liang,S.,Zheng,T., Liu, R.,Fang, H.,Tsay, S.C., & Running,S. W.(2006).Estimation ofincident

    photosynthetically active radiation from Moderate Resolution Imaging Spectrometerdata.Journal of Geophysical Research, 111, D15208. doi:10.1029/ 2005JD006730.

    Masuoka, E., Fleig, A., Wolfe, R., & Patt, F. (1998). Key characteristics of MODIS dataproducts. IEEE Transactions on Geoscience and Remote Sensing, 36(4), 13131323.

    Mlawer, E. J., Taubman, S. J., Brown, P. D., Lacono, M. J., & Clough, S. A. (1997). Radiativetransfer for in-homogeneous atmosphere: RRTM, a validated correlated-k modelfor the longwave. Journal of Geophysical Research, 102, 1666316682.

    Mu, Q., Heinsch, F. A., Zhao, M., & Running, S. W. (2007). Development of a globalevapotranspiration algorithm based on MODIS and global meteorology data. RemoteSensing of Environment, 111, 519536.

    Nishida, K., Nemani, R. R., Glassy, J. M., & Running, S. W. (2003). Development of anevapotranspiration Index Aqua/MODIS for Monitoring Surface Moisture Status.IEEE Transactions on Geoscience and Remote Sensing, 41(2), 493501.

    Nishida, K., Nemani, R. R., Running, S. W., & Glassy, J. M. (2003). An operational remotesensing algorithm of land surface evaporation. Journal of Geophysical Research, 108.doi:10.1029/2002JD002062 108, D9, 4270.

    Prata, A. J. (1996).A newlong-waveformula forestimating downward clear-sky radiation

    at the surface. Quarterly Journal Royal Meteorological Society, 122(533), 11271151.Prihodko, L., & Goward, S. (1997). Estimation of air temperature from remotely sensedsurface observations. Remote Sensing of Environment, 60, 335346.

    Reisner, J., Rasmussen, R. M., & Bruintjes, R. T. (1998). Explicit forecasting of supercooledliquidwater in winter storms usingthe MM5mesoscalemodel.Quarterly JournalRoyalMeteorological Society, 124, 10711107.

    Running, S. W., & Coughlan, J. C. (1988). A general model of forest ecosystem processesfor regional applications. I. Hydrologic balance, canopy gas exchange and primaryproduction processes. Ecological Modeling, 42, 125154.

    Ryu, Y., Kang, S., Moon, S., & Kim, J. (2008). Evaluation of Moderate Resolution ImagingSpectroradiometer (MODIS) derivedland surfaceradiation balanceover complexterrainandheterogeneous landscape forclear skydays.Agricultural and Forest Meteorology, 148(10), 15381552.

    Saigusa,N., Yamamoto,S., Hirata,R., Ohtani, Y.,Ide, R., Asanuma, J., et al. (2008).Temporaland spatial variations in the seasonal patterns of CO2 flux in boreal, temperate, andtropical forest in East Asia. Agricultural and Forest Meteorology, 148(5), 700713.

    Seemann, S. W., Li, J., Menzel, W. P., & Gumley, L. E. (2003). Operational retrieval ofatmospheric temperature, moisture, and ozone from MODIS Infrared radiations.

    Journal of Applied Meteorology, 42, 10721091.

    672 K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673

    http://dx.doi.org/10.1029/2006WR0050055http://dx.doi.org/10.1029/2006WR0050055http://dx.doi.org/10.1029/%202005JD006730http://dx.doi.org/10.1029/%202005JD006730http://dx.doi.org/10.1029/2002JD002062http://dx.doi.org/10.1029/2002JD002062http://dx.doi.org/10.1029/2002JD002062http://dx.doi.org/10.1029/%202005JD006730http://dx.doi.org/10.1029/2006WR0050055
  • 7/27/2019 Mapping Evapotranspiration Using Modis .. (1)

    17/17

    Seemann, S. W., Borbas, E. E., Li, J., Menzel, W. P., & Gumley, L. (2006). MODIS atmosphericprofile retrievalalgorithmtheoreticalbasis document(pp.134).Madison,WI: Universityof Wisconsin-Madison.

    Shin, S., & An, T. (2007). Development of estimating method for areal evapotranspi-ration using satellite data. Korean Journal of Geographic Information Studies , 10(2),7080.

    Shuttleworth, W. J.,Gurney, R. J.,Hsu, A. Y.,& Ormsby, J. P.(1989). FIFE: Thevariation inenergy partition at surface flux sites. IAHS Publication, 186, 6774.

    Stauffer, D. R., & Seaman, N. L. (1994). Multiscale four-dimensional data assimilation.Journal of Applied Meteorology, 33, 416434.

    Tang,B., Li,Z. -L.,& Zhang,R. (2006).A directmethod forestimating netsurfaceshortwave

    radiation from MODIS data. Remote sensing of Environment, 103, 115126.Venturini, V., Islam, S., & Rodriguez, L. (2008). Estimation of evaporative fraction andevapotranspiration from MODIS products using a complementary based model.Remote Sensing of Environment, 112, 132141.

    Wang, Y., Woodcock, C., Buermann, W., Stenberg, P., Voipio, P., Smolander, H., et al.(2004). Evaluationof theMODISLAI algorithm at a coniferous forest site inFinland.Remote Sensing of Environment, 91, 114127.

    Wang, K., Li, Z., & Cribb, M. (2006). Estimation of evaporative fraction from acombination of day and night land surface temperature and NDVI: A new methodto determine the PriestlyTaylor parameter. Remote Sensing of Environment, 102,293305.

    Xavier, V., Chandrasekar, A., Singh, R., & Simon, B. (2006). The impact of assimilation ofMODIS data for the prediction of a tropical low-pressure system over India using amesoscale model. International of Remote Sensing, 27(20), 46554676.

    Yamazaki, Y., & Orgaz, M. (2005). Forecasting mesoscale precipitation using the MM5model with the Four-Dimensional Data Assimilation (FDDA) technique. Global NEST

    Journal, 7(3), 258263.Zhao, M., Heinsch, F. A., Nemani, R. R., & Running, S. W. (2005). Improvements of the

    MODIS terrestrial gross and net primary production global data set. Remote Sensingof Environment, 95, 164176.

    673K. Jang et al. / Remote Sensing of Environment 114 (2010) 657673