Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

Embed Size (px)

Citation preview

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    1/18

    PLEASE SCROLL DOWN FOR ARTICLE

    This article was downloaded by: [Ingenta Content Distribution Psy Press Titles]On: 7 November 2009Access details: Access Details: [subscription number 911796918]Publisher Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,37-41 Mortimer Street, London W1T 3JH, UK

    International Journal of Remote SensingPublication details, including instructions for authors and subscription information:http://www.informaworld.com/smpp/title~content=t713722504

    Remote sensing based estimation of evapotranspiration ratesE. Boegh a; H. Soegaard aa Institute of Geography, University of Copenhagen, 1350 Copenhagen K, Denmark

    Online Publication Date: 01 July 2004

    To cite this Article Boegh, E. and Soegaard, H.(2004)'Remote sensing based estimation of evapotranspiration rates',InternationalJournal of Remote Sensing,25:13,2535 2551

    To link to this Article: DOI: 10.1080/01431160310001647975URL: http://dx.doi.org/10.1080/01431160310001647975

    Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf

    This article may be used for research, teaching and private study purposes. Any substantial orsystematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply ordistribution in any form to anyone is expressly forbidden.

    The publisher does not give any warranty express or implied or make any representation that the contentswill be complete or accurate or up to date. The accuracy of any instructions, formulae and drug dosesshould be independently verified with primary sources. The publisher shall not be liable for any loss,actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directlyor indirectly in connection with or arising out of the use of this material.

    http://www.informaworld.com/smpp/title~content=t713722504http://dx.doi.org/10.1080/01431160310001647975http://www.informaworld.com/terms-and-conditions-of-access.pdfhttp://www.informaworld.com/terms-and-conditions-of-access.pdfhttp://dx.doi.org/10.1080/01431160310001647975http://www.informaworld.com/smpp/title~content=t713722504
  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    2/18

    Remote sensing based estimation of evapotranspiration rates

    E. BOEGH* and H. SOEGAARD

    Institute of Geography, University of Copenhagen, Oester Voldgade 10,1350 Copenhagen K, Denmark

    (Received 19 June 2002; in final form 18 August 2003 )

    Abstract. A remote sensing based method is presented for calculatingevapotranspiration rates (lE) using standard meteorological field data andradiometric surface temperature recorded for bare soil, maize and wheat

    canopies in Denmark. The estimation oflE is achieved using three equations tosolve three unknowns; the atmospheric resistance (rae), the surface resistance (rs)and the vapour pressure at the surface (es) where the latter is assessed using anempirical expression. The method is applicable, without modification, to densevegetation and moist soil, but for a dry bare soil, where the effective source ofwater vapour is below the surface, the temperature of the evaporating front (Ts*)can not be represented by the measured surface temperature (Ts). In this case(Ts-Ts*) is assessed as a linear function of the difference between surfacetemperature and air temperature. The calculated lE is comparable to latentheat fluxes recorded by the eddy covariance technique.

    1. Introduction

    Spatially distributed estimates of evapotranspiration (lE) rates are significant

    both for agricultural management planning and climate studies. In agricultural

    management planning, expressions based on the atmospheric demand are nor-

    mally used to assess lE, and the physiological control of the vegetation is con-

    sidered through empirical crop coefficients (i.e. Allen et al. 1998). More precise

    solutions are usually confined to research experiments where surface fluxes and/or

    leaf stomatal resistance data are recorded.

    The PenmanMontieth (PM) equation is a well established physically based

    method for estimating evapotranspiration rate (Monteith and Unsworth 1990).Apart from standard meteorological input data, the PM equation requires

    information about surface roughness and stability conditions to calculate the

    atmospheric resistance to heat transfer (ra); and the surface resistance (rs) needs to

    be estimated. Measuring representative values of rs for vegetation canopies is very

    time-demanding, while modelling the vegetation surface (stomatal) resistance is

    difficult because plant physiological processes are controlled both by the physical

    environment and by the selective behaviour of plant species with respect to optimal

    functioning (Norman 1993). Empirical models are available for predicting the

    International Journal of Remote SensingISSN 0143-1161 print/ISSN 1366-5901 online # 2004 Taylor & Francis Ltd

    http://www.tandf.co.uk/journalsDOI: 10.1080/01431160310001647975

    *Corresponding author; e-mail: [email protected]

    INT. J. REMOTE SENSING, 10 JULY, 2004,VOL. 25, NO. 13, 25352551

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    3/18

    stomatal resistance based on climatic and soil moisture input data (i.e. Jarvis 1976,

    Lindroth and Halldin 1986), but they are not generally applicable without prior

    local determination of the empirical constants. Instead, semi-empirical models

    based on functional relationships between photosynthesis and stomatal resistance at

    the leaf scale (Ball et al. 1987, Leuning 1995) may be preferred. In this case, the

    evapotranspiration/rs-model is coupled to a mechanistic photosynthesis-model (i.e.

    Collatz et al. 1991, Sellers et al. 1992, Leuning 1995, Boegh et al. 1999, Thorgeirsson

    and Soegaard 1999, Boegh et al. 2002b, among many others). Even though the

    process of computing the evapotranspiration rate becomes more complex, the

    process of fitting empirical stomata models to a given location and situation is

    substituted by functional parameters related to vegetation type.

    When the surface is only partly covered by vegetation, the estimation of

    evapotranspiration is further complicated because net radiation needs to be

    partitioned between soil and vegetation, and a more complex two-layer model

    structure (i.e. Shuttleworth and Wallace 1985) is required to allow for the exchange

    of energy between soil and vegetation (i.e. within-row advection). From an

    operational standpoint, two-layer models are particularly difficult to apply because

    of the additional need to estimate the surface resistance of the soil. In practice,

    empirical models relating the soil resistance to soil surface moisture do not give

    good estimates of hourly and daily evaporation (Daamen and Simmonds 1996), and

    alternative methods may be needed for estimating soil evaporation rates (Ham and

    Heilman 1991, Yamanaka et al. 1997, van der Keur et al. 2001). Theoretically, the

    application of the PM equation (to one or two layers) is not well suited for

    calculating evaporation rates from dry bare soils because its derivation is based on

    the assumption that the sources of water vapour and heat are represented at the

    same temperature, Ts (i.e. Monteith and Unsworth 1990). In reality, the tem-

    perature of a dry bare soil surface may be much higher than the temperature of the

    evaporating front (Ts*) which is located deeper in the soil.

    Nevertheless, if remote sensing data of Ts are available, estimating the bulk

    stomatal resistance, the soil surface resistance (for bare soil/sparse vegetation) and

    the surface humidity (of dry soils) can be avoided. In such cases, Ts is taken to

    represent the aerodynamic temperature at canopy source height (T0) to enable the

    calculation of the sensible heat flux (H):

    H~rcp T0{Ta

    ra W m{2

    1

    where rcp is the heat capacity (J m23 C21). The evapotranspiration rate can thenbe assessed as the residual in the surface energy balance equation (equation (2))

    after estimating or measuring net radiation and soil heat flux (i.e. Seguin et al. 1989,

    Humes et al. 1997). However, two significant drawbacks of this method are that Tsmay exceed T0, even for dense canopies (Choudhury et al. 1986, Huband and

    Monteith 1986, Kalma and Jupp 1990) and that the atmospheric resistance between

    the surface and the air (rae) may be larger than the atmospheric resistance between

    canopy source height and the air above the canopy (ra) (i.e. Stewart et al. 1994).

    Numerous empirical and theoretical studies have been conducted to determine an

    excess resistance to heat transfer between the surface and source height (i.e.

    Kustas et al. 1989, Sugita and Brutsaert 1990, Stewart et al. 1994, Troufleau et al.1997, Verhoef et al. 1997, Massman 1999). The classical approach is to use the

    quantity kB21 to estimate the excess resistance, rex~kB21/(ku.), where k is the von

    Karman constant, u. is the friction velocity (m s21), and kB21 is the natural

    logarithm of the ratio of the roughness length for momentum (z0) and roughness

    2536 E. Boegh and H. Soegaard

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    4/18

    length for heat (z0h). In remote sensing studies, the kB21 is normally determined

    empirically. For dense canopies, it is approximately constant (around 2), but for

    sparse canopies, it is usually higher (Kustas et al. 1989, Stewart et al. 1994,

    Troufleau et al. 1997, Verhoef et al. 1997), and its dependence on both vegetation

    structure, climate and surface conditions has been demonstrated (Troufleau et al.1997). Very complex models are needed to predict this kB21 dependency (Massman

    1999), and currently alternative methods are being developed to estimate lE from

    Ts where the estimation of the excess resistance to heat transfer is not needed.

    Instead, simultaneous measurements of Ts from two different view angles can be

    used to calculate lE (Norman et al. 1995, Kustas and Norman 1997) or, assuming

    the atmospheric resistance of dry and wet soils to be similar, the surface

    temperature of a dry soil column can be included for calculating soil evaporation

    rates (Qiu et al. 1998). Methods to assess the atmospheric resistance were suggested

    by Bastiaanssen et al. (1998) who used the slope of the Ts-albedo relationship of dry

    surfaces to assess the atmospheric resistance from Earth observations, and Boeghet al. (2002a) who introduced an empirical equation for the vapour pressure at the

    surface (es) to facilitate mathematical solutions for rae and rs. Because the method

    of Boegh et al. (2002a) used the measurement of Ts to assess the surface saturation

    vapour pressure, this method needed adjustment to assess the surface humidity (and

    lE) for dry bare soils where the effective source area of water vapour is below the

    surface.

    In the present paper, the method of Boegh et al. (2002a) is further elaborated

    and expanded to include an adjustment function accounting for the difference in

    temperature between the surface and the evaporating front (Ts-Ts*) of dry soil

    surfaces. The method requires inputs of surface temperature, net radiation, soil heatflux, air temperature and vapour pressure. Because the latter two variables are

    standard meteorological data and the remaining inputs can be obtained from

    remote sensing data, the method could be of particular relevance for monitoring

    spatially distributed evapotranspiration using satellite data such as Landsat,

    NOAA-AVHRR or EOS-MODIS. In the present paper, the utility of the metho-

    dology is demonstrated using thermal infrared field data which were recorded for

    bare, sparsely and densely vegetated soils in Denmark. In Boegh et al. (2004), the

    method is applied to AVHRR observations and meteorological data which were

    predicted by a high resolution weather forecast model.

    2. Methodology

    The evapotranspiration rate (lE) relates to the surface energy balance as

    lE~Rn{H{G W m{2

    2

    where Rn is net radiation (W m22) and G is the soil heat flux (W m22). In

    equation (2), H and lE may be expressed as

    H~rcp Ts{Ta

    rae W m{2

    3

    lE~ rcpc

    es{ea =rae W m

    {2 4

    where Ts is the surface temperature (C), c is the psychrometer constant (Pa C21),

    ea is the vapour pressure above the canopy (Pa), es is the vapour pressure at the

    surface (Pa), and rae is the atmospheric resistance to heat transfer between the

    surface and the air above the canopy (s m21), which is given by rae~razrex.

    Estimation of evapotranspiration rates 2537

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    5/18

    Assuming the atmospheric resistance to be the same for temperature and water

    vapour, an expression for rae is obtained by combining equations (2), (3) and (4)

    rae~rcpTs{Ta z es{ea =c

    Rn{Gs m{1

    5

    An equation for the surface resistance (rs) is obtained on the basis of the followingequations; lE~ rcp

    c

    es{es

    rs~ rcp

    c

    es{ea =rae, where the saturated

    vapour pressure at the evaporating front (es*) is evaluated at the temperature

    (Ts*). By rearrangement, rs is given by

    rs~rae es{es

    es{ea s m

    {1

    6

    For dense canopies and moist soils, Ts* equals Ts, but for a dry bare soil, the

    evaporating front is located beneath a dry surface layer exhibiting different thermal

    properties, and Ts may be much higher than Ts* in this situation. Because an

    increase in Ts due to greater solar radiation increases both the soil and the

    atmospheric sensible heat flux (Garratt 1992, Cellier et al. 1996), the differencebetween Ts and Ts* may be inferred from the observations of (Ts-Ta). The close

    coupling of G and H through the surface temperature allows G to be expressed as a

    fraction of H, i.e. G~aH (Berkowicz and Prahm 1982, Novak 1986). Since

    G~ks Ts{Ts

    zd (ks is thermal conductivity of dry soil and zd is the thickness of

    the upper dry soil layer), insertion in equation (3) leads to the following expression

    for (Ts-Ts*)

    Ts{Ts

    ~B Ts{Ta

    0C 7

    where B~arcpzd/(raeks). Equation (7) implies the use of the depth of the evaporating

    surface instead of the more traditional use of surface-moisture availability forcalculating soil-limited evaporation rates. A simple energy-balance model applying

    zd instead of surface-moisture for calculating lE was also proposed by Yamanaka

    et al. (1997). In this study, the coefficient B is evaluated using field data and

    literature values of zd and ks of the dry soil layer (4.3).

    To solve rae (equation (5)) and rs (equation (6)), an equation for the third

    unknown, es, is required. Because es is regulated at the surface by the amount of

    feedback which is mediated both by the atmospheric boundary layer and the

    surface response functions, its evaluation should take place within a larger scale

    environment, such as that represented by the PM equation (Jones 1992; p. 117). By

    rewriting the PM equation, the vapour pressure deficit at the surface is given by(Jarvis and McNaughton 1986):

    es{es

    ~V es{es

    eqz 1{V ea{ea

    8

    where V is the decoupling coefficient, V~(D/cz1)/(D/cz1zrs/rae), which quantifies

    the degree of atmospheric coupling between the surface and the air; and D is

    the slope of the saturation vapour pressure versus temperature relationship

    which is evaluated at (TszTa)/2. With an efficient vertical mixing of the air,

    rae decreases and Vp0 in which case the atmospheresurface coupling is good.

    If the surface is completely coupled to the atmosphere (V~0), the vapour

    pressure deficit of the atmosphere (ea*-ea) is imposed at the surface (equation (8)).The surface vapour pressure deficit during decoupled conditions, (es*-es)eq, is

    solved by es{es

    eq~lEeqrs c

    rcp

    , where lEeq~D(Rn-G)/(Dzc) is the equilibrium

    evapotranspiration rate which is approached during decoupled conditions (Vp1)

    (Jarvis and McNaughton 1986).

    2538 E. Boegh and H. Soegaard

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    6/18

    When (Ts-Ta) increases above 510C, the linear approximation ofD in the PM

    equation is invalid (Paw U and Gao 1988). This temperature difference is typically

    reached for bare dry soils in Denmark. When nonlinear solutions are used for the

    saturated vapour pressure function instead of the linear approximation, it can be

    shown (Paw U and Gao 1988) that the limit of lE during decoupled conditionsapproaches the available energy (Rn-G) which may be very different from the lEeqused to calculate (es*-es)eq in equation (8). The application of equation (8) for

    evaluation of es is therefore in error when rs/raep0. In addition, it was already

    discussed that differences between Ts and Ts* may be substantial for bare dry soils

    which further invalidates the applicability of PM-theory in such situations.

    In the present study, measurements rather than predictions are used to assess the

    limit of es at any time. If the surface is decoupled from the atmosphere, the

    atmospheric vertical transfer of vapour is weak. In this case, water vapour tends to

    accumulate at the surface until, eventually, it approaches es*. The limit of es during

    decoupled conditions may therefore be calculated from the Ts measurement, or theTs* estimate (for a non-saturated surface). On the contrary, if the surface is well

    coupled to the atmosphere (efficient vertical mixing of air), es is expected to

    approach ea. In this paper, the feasibility of V as a weighting factor to place esbetween its (measurement-based) limit values, es* and ea, is investigated, i.e.

    es~A V esz 1{V ea Pa 9

    Substituting equation (9) in equation (6) it can be seen that the empirical factor A

    appears to be analytically related to ea/es* by

    A~craezcrszDraezDrea

    es

    craezcrszDraezDrs10

    For ea/es*~1, or rs~0, then A~1. For lower values of ea/es*, Av1. In this paper, a

    constant value of A (A~0.9) in equation (9) facilitates the calculation of

    evapotranspiration rates during a wide range of conditions. For this purpose,

    the hypothetical expression for es (equation (9) using A~0.9) is solved together with

    equations (5) and (6) using NewtonRaphson iteration (appendix A), thereby

    allowing the derived rae and rs to be used for computing lE:

    lE~ rcpc

    es{ea

    raezrs W m

    {2

    11

    Apart from evaluating the usefulness of the simple empirical expression for es by

    comparing the ensuing calculations of rae and lE with measurement-based values ofrae (rae~rcp(Ts-Ta)/H) and eddy covariance of lE data, the dependency of es on V

    is also evaluated relative to the results given by a more complex canopy scale

    vegetationatmosphere transfer (VAT) model. The VAT model applied in this

    paper is a slightly simplified version of the model described in Boegh et al. (2002b).

    It provides an iterative solution for surface fluxes and surface conditions within a

    larger scale environment, and it is validated using the field measurements of lE and

    Ts. A summary of the VAT model is given in appendix B.

    3. Location and equipmentFor the present study, data collected from two fields cropped with maize

    and winter wheat, respectively, are used. The fields are located in an agricultural

    area in Denmark (9.423 E, 56.486 N) characterized by sandy loam soils, and the

    measurement campaign took place during the spring and summer of 1998.

    Estimation of evapotranspiration rates 2539

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    7/18

    Generally, the growth season in 1998 was very humid (figure 1(a)). The winter

    wheat was sown in Autumn 1997, and the maize was sown in May 1998. During the

    measurement campaign, the leaf area index ranged from 0 to 4.2 for winter wheat,

    and from 0 to 4.4 for maize (figure 1(b)).

    Leaf area index was measured each week using destructive sampling (Boeghet al. 2002b). A continuous record of the field surface temperature was made in

    1998, except for the period 24 July to 6 August for wheat, and in the period 4

    August to 9 August for maize. On each field, thermal infrared radiometers (KT17,

    Heimann, D; Wiesbaden, Germany) were mounted on masts at 2 m height with

    azimuth angles of 60 and pointing 45 towards the surface. An emissivity of 0.99

    was applied to convert the thermal radiance to surface temperature. Net radiation

    (Q*6, REBS Inc., USA), soil temperature, air temperature and air humidity

    (MP300, Rotronic; Bassersdorf, Switzerland) were measured continuously through-

    out the study period.

    Sensible and latent heat fluxes were measured using the eddy covariancetechnique (Baldocchi et al. 1988). The winter wheat field size is 600 m6800 m, and

    the minimum fetch is 300 m. The maize field, which is long and narrow, is located in

    an eastwest direction. For this field, data were abandoned in the period between

    day 151 and 196 due to an insufficient fetch (Soegaard et al. 2002). The equipment

    at each field consisted of a 3-dimensional sonic anemometer (Gill Solent;

    Lymington, UK), and an IRGA (Infrared Gas Analyzer; Li-Cor, USA) was

    used for measuring water vapour concentrations. The sampling tube (Bev-a-line, Li-

    Cor; Lincoln, NE, USA) had an inner diameter of 3.2 mm and it was 10 m in

    length. The sensor head was mounted 2.5 m above the top of the canopy. The air

    was sucked through the sampling tube by a membrane pump at a flow rate of9lmin21. The Edisol software package was used for data collection and processing

    (Moncrieff et al. 1997). The flux data availability is indicated in figure 1(b).

    4. Results

    4.1. The dependency of es on V for dense vegetation

    Because the stomatal cavities of leaves are usually close to saturation, it can be

    assumed that Ts~Ts* for dense vegetation canopies. In order to study the

    dependency of es on V in such situations, the VAT model (appendix B) is applied to

    Figure 1. (a) Mean midday temperature (11.0014.00 h) and precipitation in the experi-mental period. (b) Leaf area index (LAI) of wheat (empty circle) and maize (filledcircle) in the experimental period. The horizontal lines represent the flux dataavailability for wheat (upper line) and maize (lower line).

    2540 E. Boegh and H. Soegaard

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    8/18

    evaluate the es-V relationship of a wheat canopy in periods with a dense vegetation

    cover (L~4). For inclusion of variability in the environment, the VAT model is

    applied both to a dry period (1821 May 1998) and a wet period (2730 May

    1998). The chosen dry period is towards the end of the longest dry period in the

    study-area that year (the last rain fell on 12 May 1998), whereas the wet periodfollows the occurrence of heavy rainfall (17 mm) on 26 May 1998 (figure 1(a)).

    The simulated fluxes and surface conditions by the VAT model are in quite

    good agreement with the field measurements of lE and Ts for wheat (figure 2).

    Assuming that the model is also capable of predicting es and rs correctly (rae is

    calculated from measurements; appendix B), the VAT model can be used to study

    the dependency of es on V.

    As would be expected, the measurements of ea at reference height are

    independent of variations in V (figure 3(a)). The surfaceatmosphere system is well

    coupled (Vp0) during early mornings and late afternoons because the leaf stomatal

    (surface) resistance is higher during these periods. In such situations, the simulatedes (figure 3(b)) remains in the same range as ea. Towards the less well coupled

    conditions prevailing during midday (Vp1), es increases (figure 3(b)), and the same

    pattern is recognized in the relationship between es* and V (figure 3(c)). While es is,

    in fact, linearly related to es*, it is even better related to the results derived from the

    empirical expression for es suggested by equation (9) (figure 4). Using the VAT

    model for calculating the factor A (equation (10)), its mean value is found to be 0.87

    (SD~0.06) during the 6 days depicted in figures 2 and 3.

    4.2. lE calculation based on empirical expression for es

    The calculations to estimate lE (equations (5), (6), (9) and (11)) were run on ahalf-hourly basis using the inputs of Rn, Ts, Ta, ea and G during the periods when

    flux data are available for validation (figure 1(b)). For this purpose, G is determined

    as the residual in the energy balance equation (equation (2)) using the input data of

    lE, H and Rn. At first, the calculations of lE are conducted assuming Ts~Ts*

    which is valid for dense canopies. In this case, it is seen (figure 5) that the lE

    predictions compare well with the measurements when LAI>3 (dense vegetation).

    Figure 2. (a) Comparison between the evapotranspiration rates (lE) simulated by avegetationatmosphere transfer (VAT) model (appendix B) and field measurements oflE. (b) Comparison of surface temperature (Ts) simulated by a VAT model and fieldmeasurements of Ts. The figures were produced on the basis of data representingdense vegetation, including both a dry period (18 May to 21 May) and a wetperiod (27 May to 30 May, 1998).

    Estimation of evapotranspiration rates 2541

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    9/18

    Figure 3. (a) Measurements of atmospheric vapour pressure (ea) versus the decouplingcoefficient (V) simulated by a vegetation-atmosphere transfer (VAT) model, (b) VATsimulations of the surface vapour pressure (es) versus V, (c) VAT simulations of thesaturated vapour pressure at the surface (es*) versus V. The figures were produced onthe basis of data representing dense vegetation, including both a dry period (18 Mayto 21 May, 1998) and a wet period (27 May to 30 May, 1998).

    Figure 4. Comparison between the vapour pressure at the surface (es) estimated using thehypothetical equation (8), es~AVe

    sz 1{V ea, and es modelled by a more complex

    vegetationatmosphere transfer model. (a) using A~1 in equation(9), (b) usingA~0.9 in equation (9). The figures were produced on the basis of data representingdense vegetation, including both a dry period (18 May to 21 May, 1998) and a wetperiod (27 May to 30 May, 1998).

    2542 E. Boegh and H. Soegaard

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    10/18

    On many occasions, the results also perform well for LAIv3, in particular for

    wheat. The better performance for the wheat field than for maize relates to

    differences in (Ts-Ta).

    Generally, the larger values of (Ts-Ta) predominate at low LAI, but these are

    much lower for wheat (figure 6(a)) than for maize (figure 6(b)) because the periods

    with sparse vegetation at the wheat field were characterized by frequent rain events

    (figure 1(a)). In contrast, the exposed soil patches at the maize field cause (Ts-Ta)

    to increase dramatically during the dry spells in MayJune as can be seen in

    figure 1(b)).Overall, the lE predictions (assuming Ts~Ts*) are in good accordance with field

    measurements when (Ts-Ta)v5C (figure 7). However, when (Ts-Ta)w5C, the

    calculated lE is overestimated. In this situation, the effective source of water

    vapour from the dry soil patches at the maize field is located below an upper dry

    layer exhibiting distinct thermal properties. This causes the measurement of Ts to

    exceed the hypothetical temperature of the evaporating front which is located

    deeper in the soil (Ts*) which, in turn, causes es* and lE to be overestimated.

    Figure 5. Comparison between evapotranspiration rates (lE) calculated assuming Ts~Ts*(using equations (5), (6), (8) and (9)) and eddy covariance data of lE. Results areshown separately for bare/sparse canopies where LAIv3 (empty circle) and for densecanopies where LAI>3 (filled circle). (a) Wheat, (b) maize.

    Figure 6. Field data of surface temperature (Ts) and air temperature (Ta) are used toillustrate the variations in (Ts-Ta) for (a) wheat and (b) maize during the experimentalperiod.

    Estimation of evapotranspiration rates 2543

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    11/18

    4.3. Assessment of Ts* of dry bare soil

    The estimation of Ts* during dry conditions can be made using equation (7)

    if information on zd is available. Even though the moisture condition in the

    upper thin surface layer of soils is generally considered to be very important for

    determining rs and soil-water limited rates of lE, only very few registrations of zdhave been reported in the literature. For a sandy loam soil in Japan with a particle

    composition (81% sand, 9% silt, 10% clay) very similar to that of the agricultural

    topsoils in the present study (80% sand, 10% silt, 10% clay), the maximum depth of

    the dry layer was found to be 0.02 m while for a sand dune site (96% sand, 2% silt,

    2% clay), the maximum observed zd was 0.08 cm (Yamanaka and Yonetani 1999).

    Using zd~0.02 m to represent the soil condition during the very dry period,

    1821 May 1998, in this study, measurement-based estimates of G/H (equation (2))

    and rae (equation (3)) can be used to calculate B. In this period, the average G/H

    was found to be 0.66 (SD~0.29) and the average rae was 53sm21 (SD~1 3 s m21).

    Using ks~0.25Wm21

    C21 to represent the thermal conductivity of the upper dry

    soil layer, B is given by B~(0.66*1227*0.02)/(53*0.25)~1.2. Theoretically, the

    factor B should then vary from 0 (in which case Ts~Ts*) for a wet soil to 1.2 for a

    dry soil. Because the transition from energy-limited to soil-limited evaporationtends to be abrupt and is accompanied by an increase in surface temperature

    (Amano and Salvucci 1999), it is simply assumed that Ts*5C. The sensitivity of the calculated lE to B is evaluated in 4.4.

    4.4. lE calculation using empirical estimates of es and Ts*

    In the following, the calculation of lE for surfaces characterized by (Ts-

    Ta)v5C is based on the assumption that Ts~Ts* while the calculation of lE for

    warm surfaces with (Ts-Ta)>5C uses the relationship in equation (7) to assess Ts*.

    This causes predictions to be comparable with measurements (figure 8).Figure 9 facilitates a more detailed evaluation of the results at the wheat field

    and maize field, respectively. For each field, diurnal predictions of lE and rae are

    compared with measurement-based values of lE and rae on three different days

    varying in vegetation density and surface temperature (table 1).

    Figure 7. Comparison between evapotranspiration rates (lE) calculated assuming Ts~Ts*(using equations (5), (6), (8) and (9)) and eddy covariance data of lE. Results areshown separately for (Ts-Ta)w5C (empty circle) and for (Ts-Ta)5C (filled circle),where (Ts-Ta) is the difference between surface temperature and air temperature.

    (a) Wheat, (b) maize.

    2544 E. Boegh and H. Soegaard

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    12/18

    Towards the mornings and late evenings, the predictions of rae obtain

    large (positive or negative) values when the denominator in equation (5) (Rn-G)

    approaches zero. The case is the same for the measurement-based rae when H

    approaches zero. During midday, the predictions of rae are in quite good agreement

    with the measurement-based rae. For the dense maize field (figure 9( f)), Ts typically

    comes very close to Ta during afternoons, and it may even happen that Ts becomes

    lower than Ta. Simultaneous observations of negative (Ts-Ta) and positive H

    indicate the presence of counter-gradient fluxes in the tall maize canopy which

    causes the measurement-based rae to become negative (figure 9( f)). This situationillegitimates both the measurement-based rae and the applied methodology for

    calculating lE. On day 222 (figure 9( f)), both (Ts-Ta) and H turn negative during

    late afternoon which indicates that sensible heat is being used to increase the

    transpiration rates.

    The sensitivity of the calculated lE in figure 9 to the various inputs and

    parameter settings is illustrated in table 2. Overall, there is a quite broad range of

    sensitivities to the individual parameters. For sparse/dense canopies with (Ts-

    Ta)v5C, the sensitivity is highest to the observations of Ts, Rn and ea while the

    sensitivity to the choice of parameter values for A and B is less. For the dry bare

    soil and sparsely vegetated surfaces where (Ts-Ta)>5C, the sensitivity to theempirical surface humidity parameter A is highest, and the sensitivity to the

    parameter B is also important. Nevertheless, it seems feasible to use a fixed

    parameter value, A~0.9, regardless of vegetation density and dryness condition

    (figure 8). Optimally, the setting of parameter B should use information on the dry

    surface layer development (4.3) to improve the predicted transition from energy

    limited to soil water limited evapotranspiration rates. However, in this study

    reasonable results were obtained using a setting of B~1.2 to represent a sandy

    loam soil.

    5. ConclusionA method is presented to calculate lE using standard meteorological data and

    remote sensing based estimates of Ts, Rn and G. The method uses physical

    descriptions of the surface energy exchange together with an empirical assessment

    of the surface humidity (equation (9)) and, for dry surfaces, the temperature of the

    Figure 8. Comparison between predicted evapotranspiration rates (lE) and field measure-ments of lE for (a) wheat and (b) maize. The 1:1 lines are also shown. Thecomparison includes results from all days with available flux data (figure 1(b)).

    Estimation of evapotranspiration rates 2545

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    13/18

    evaporating front needs to be evaluated (equation (7)). It is interesting that, whenwe use a more basic and traditional, although more complicated, model, we find

    that the hypothetical equation (9) is approximately true. For humid surface

    conditions (Ts*

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    14/18

    boundary condition for the atmospheric fluxes and the upper boundary condition

    for the soil fluxes (Ts-Ts*) and (Ts-Ta) are closely associated. Theoretically, this

    relationship is dependent on the depth of the evaporating surface, but in practice, a

    constant coefficient representing the ratio of (Ts-Ts*) and (Ts-Ta) for a dry soil

    was found useful for estimating Ts* and calculation of lE for dry surfaces with

    (Ts-Ta)>

    5

    C.Should the expressions for es and Ts* prove to be reasonable approximations

    in other studies, they may provide an effective and simple way to calculate

    evapotranspiration from standard, easily obtainable meteorological data and

    thermal infrared remote sensing data. In Boegh et al. (2004), the application of the

    method for calculating lE in Denmark is demonstrated based on inputs of Earth

    observations (NOAA AVHRR) and climate predictions which were computed by a

    high-resolution weather forecast model.

    Appendix A

    The NewtonRaphson Method states that if x~c is an approximation to the

    solution of the equation f(x)~0, then a better approximation is given by

    cnew~c{f c

    f0 c A1

    Table 1. Leaf area index, LAI, and the difference between surface temperature and airtemperature at 13.00 h (Ts-Ta)13h, on selected days for (a) wheat and (b) maize. Seefigure 9 for results of evapotranspiration rates and atmospheric resistance on thesedays.

    (a) Wheat

    Day 112 Day 138 Day 189LAI 0.3 3.8 0(Ts-Ta)13h 4.4C 3.6C 4.5C

    (b) Maize

    Day 138 Day 164 Day 222LAI 0 0.1 4.2(Ts-Ta)13h 15.1 6.4C 2.1C

    Table 2. Fractional change of the calculated evapotranspiration rate to independentincreases of 10% in input and parameter settings. Results are shown for wheat andmaize on selected days at midday (13.00 h). See also table 1 and figure 9. (Ts, surfacetemperature; Rn, net radiation; Ta, air temperature; ea, vapour pressure; A, parameterin equation (9); B, parameter in equation (7).

    Wheat Maize

    Day 112 Day 138 Day 189 Day 138 Day 164 Day 222

    Ts 20.57 20.45 0.26 20.26 20.17 0.14Rn 0.09 0.08 20.63 0.09 0.11 0.09T

    a0.05 0.07 0.07 0.26 0.01 0.13

    ea 20.05 20.06 20.03 0.01 20.31 0.33A 0.08 0.07 0.03 0.39 0.43 0.02B 0 0 0 20.14 20.18 0

    Estimation of evapotranspiration rates 2547

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    15/18

    When estimating lE, the function f is given by

    f x ~rnewae {rae A2

    where rnewae is calculated using equation (5). The derivative of f follows as

    f0 x ~drnew

    aedre{1 A3

    where

    drnewaedre~

    drnewaedes

    des

    dreA4

    with

    drnewaedes~

    rcp

    c Rn{G A5

    and

    des

    drae~es A

    D=cz1

    D=cz1zrs=rae 2

    rs

    r2ae{D=cz1

    D=cz1zrs=rae 2

    rs

    r2aeea A6

    Appendix B. Description of VAT model

    The VAT-model consists of a coupled photosynthesis/stomatal resistance model

    which is connected to the above atmosphere by an atmospheric boundary layer.

    The canopy photosynthetic rate is simulated to allow the prediction of stomatal

    (surface) resistance which, in turn, is used to evaluate the transpiration rate and the

    surface energy balance. The model is applied to a dense canopy only. The model

    applied in this paper is a slightly simplified version of the model described in Boeghet al . (2002b). The same model parameters representing wheat (taken from

    the literature) are applied, but in this paper, more input data are based on

    measurements (rather than predictions) in order to maximize the accuracy of the

    simulations. As such, inputs of measured net radiation are used in this version

    (rather than global radiation), and the atmospheric resistance between the surface

    and the atmosphere above the canopy (rae) is assessed using the measurements ofH,

    Ts and Ta:

    rae~rcp Ts{Ta

    H s m{1

    B1

    In Boegh et al. (2002b), all leaf level calculations were made separately forshaded and sunlit leaves, but in this version, only the photosynthetic rate is

    computed seperately for sunlit and shaded leaves. The radiation absorbed by sunlit

    and shaded leaves is evaluated, and an exponential profile of leaf Rubisco capacity

    in the canopy is distributed between sunlit and shaded leaves. The canopy

    photosynthesis (Ac) is the sum of shaded and sunlit fractions. The (big-leaf)

    stomatal conductance (gs~1/rs) is then calculated as (Leuning 1995):

    gs~gs0 Lza wsf Ac P

    ps{C 1zDs=D0 mol m{2s{1

    B2

    where P is atmospheric pressure (Pa), ps is the CO2 partial pressure at the surface

    (Pa), C is the CO2 compensation point (Pa), Ds is the vapour pressure deficit at thesurface, and gs0, a and D0 are empirical coefficients for wheat. The wsf is a water-

    stress factor (set to unity). Because Ds is not known, Ds and gs are solved by

    iteration. In Boegh et al. (2002b), the net-isothermal form of the PM equation was

    used to compute the transpiration rates, but in this version, it is preferable to keep

    2548 E. Boegh and H. Soegaard

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    16/18

    the VAT model completely independent of the linearly approximated definition of

    D which is used in the empirical expression for es (to be compared with the VAT-

    modelled es). The big-leaf transpiration rate is therefore computed by

    lE~ rcpc

    es{ea

    raezrs W m

    {2

    B3

    and the big-leaf surface temperature (Tl) is calculated as

    Tl~Hraercp

    zTa

    0C B4

    where H is the sensible heat flux (H~Rn-lE). The evaluation of the big-leaf

    temperature, rather than a separate evaluation of energy balances for shaded and

    sunlit leaves (as in Boegh et al. 2002b), allows direct validation of the modelled Tlrelative to the Ts measurements.

    Finally, es is calculated by

    es~es{lE rs c

    rcp

    0C B5

    The gs, Tl, lE and Ac are solved simultaneously with the computation of es, ps andpi (leaf internal CO2 pressure) through utilizing a NewtonRaphson iterative

    solution process for pi, as introduced by Collatz et al. (1991). In order to simulate

    a dense winter wheat canopy, the leaf area index is set to 4, and the leaf

    photosynthetic capacity at the top of the canopy, Vmax0, is set to 150 mmolm22s21

    to represent wheat (Wang and Leuning 1998).

    Acknowledgments

    The paper was prepared within the projects EO-flux-budget and RS-model

    which is financed by the Danish research council, Copenhagen. Information on leaf

    area index and precipitation was kindly provided by Anton Thomsen and TomJensen from The Danish Institute of Agricultural Sciences in Foulum, Denmark.

    ReferencesALLEN, R., PEREIRA, L. S., RAES, D., and SMITH, M., 1998, Crop evapotranspiration.

    Irrigation and Drainage Paper 56 (Rome, Italy: FAO).BALDOCCHI, D. D., HICKS, B. B., and MEYERS, T. P., 1988, Measuring biosphere

    atmosphere exchanges of biologically related gases with micrometeorologicalmethods. Ecology, 69, 13311340.

    BALL, J . T. , WOODROW, I . E., and BERRY, J. A., 1987, A model predicting stomatalconductance and its contributors to the control of photosynthesis under differentenvironmental conditions. In Progress in Photosynthesis Research, Vol. IV, edited by

    J. Biggens (Dordrecht: Martinus Nijhoff), pp. 221225.BASTIAANSSEN, W. G. M., MENENTI, M., FEDDES, R. A., and HOLTSLAG, A. A. M., 1998, A

    remote sensing surface energy balance algorithm for land (SEBAL). 1. Formulation.Journal of Hydrology, 212213, 198212.

    BERKOWICZ, R., and PRAHM, L. P., 1982, Sensible heat flux estimated from routinemeteorological data. Journal of Applied Meteorology, 21, 18451864.

    BOEGH, E., SOEGAARD, H., FRIBORG, T., and LEVY, P. E., 1999, Models of CO2 and watervapour fluxes from a sparse millet crop in the Sahel. Agricultural and ForestMeteorology, 93, 726.

    BOEGH, E., SOEGAARD, H., and THOMSEN, A., 2002a, Evaluating evapotranspiration ratesand surface conditions using Landsat TM to estimate atmospheric resistance andsurface resistance. Remote Sensing of Environment, 79, 329343.

    BOEGH, E., SOEGAARD, H., BROGE, H., HASAGER, C. B., JENSEN, N. O., SCHELDE, K., andTHOMSEN, A., 2002b, Airborne multispectral data for quantifying leaf area index,nitrogen concentration, and photosynthetic efficiency in agriculture. Remote Sensingof Environment, 81, 179193.

    BOEGH, E., SOEGAARD, H., CHRISTENSEN, J. H., HASAGER, C. B., JENSEN, N. O., NIELSEN,N. W., and RASMUSSEN, M. S., 2004, Combining weather prediction and remote

    Estimation of evapotranspiration rates 2549

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    17/18

    sensing data for the calculation of evapotranspiration rates: application to Denmark.International Journal of Remote Sensing, in press.

    CELLIER, P., RICHARD, G., and ROBIN, P., 1996, Partition of sensible heat fluxes into baresoil and the atmosphere. Agricultural and Forest Meteorology, 82, 245265.

    CHOUDHURY, B. J . , REGINATO, R. J., and IDSO, S. B., 1986, An analysis of infraredtemperature observations over wheat and calculation of latent heat flux. Agriculturaland Forest Meteorology, 37, 7588.

    COLLATZ, G. J., BALL, J . T. , GRIVET, C., and BERRY, J. A., 1991, Physiological andenvironmental regulation of stomatal conductance, photosynthesis and transpiration:a model that includes a laminar boundary layer. Agricultural and Forest Meteorology,54, 107136.

    DAAMEN, C. C., and SIMMONDS, L. P., 1996, Measurement of evaporation from bare soiland its estimation using surface resistance. Water Resources Research, 32, 13931402.

    GARRATT, J. R., 1992, Extreme maximum land surface temperatures. Journal of AppliedMeteorology, 31, 10961105.

    HAM, J. M., and HEILMAN, J. L., 1991, Aerodynamic and surface resistance affecting energytransport in sparse crops. Agricultural and Forest Meteorology, 53, 267284.

    HUBAND, N. D. S., and MONTEITH, J. L., 1986, Radiative surface temperature and energybalance of a wheat canopy. 1. Comparison of radiative and aerodynamic tem-perature. Boundary-Layer Meteorology, 36, 117.

    HUMES, K. S., KUSTAS, W. P., and GOODRICH, D. C., 1997, Spatially distributed sensibleheat flux over a semi-arid watershed. Part 1: Use of radiometric surface temperaturesand a spatially uniform resistance. Journal of Applied Meteorology, 36, 281293.

    JARVIS, P. G., 1976, The interpretations of the variation in leaf water potential and stomatalconductance found in canopies in the field. Philosophical Transactions of the RoyalSociety of London, Series B, 273, 593610.

    JARVIS, P. G., and MCNAUGHTON, K. G., 1986, Stomatal control of transpiration. Scalingup from leaf to region. Advances in Ecological Research, 15, 149.

    JONES, H. G., 1992, Plants and Microclimate, 2nd edn (Cambridge: Cambridge University

    Press).KALMA, J. D., and JUPP, D. L. B., 1990, Estimating evaporation from pasture using infraredthermometry: evaluation of a one-layer resistance model. Agricultural and ForestMeteorology, 51, 223246.

    KUSTAS, W. P., CHOUDHURY, B. J., MORAN, M. S., REGINATO, R. J., JACKSON, R. D., GAY,L. W., and WEAVER, H. L., 1989, Determination of sensible heat flux over sparsecanopy using thermal infrared data. Agricultural and Forest Meteorology, 44,197216.

    KUSTAS, W. P., and NORMAN, J. M., 1997, A two-source approach for estimating turbulentfluxes using multiple angle thermal infrared observations. Water Resources Research,33, 14951508.

    LEUNING, R., 1995, A critical appraisal of a combined stomatal-photosynthesis model for C,

    plants. Plant, Cell Environment, 18, 339355.LINDROTH, A., and HALLDIN, S., 1986, Numerical analysis of pine forest evaporation and

    surface resistance. Agricultural and Forest Meteorology, 38, 5979.MASSMAN, W. J., 1999, A model study of kB{1H for vegetated surfaces using localized near-

    field Lagrangian theory. Journal of Hydrology, 223, 2743.MONTEITH, J. L., and UNSWORTH, M. H., 1990, Principles of Environmental Physics, 2nd edn

    (New York: Edward Arnold).NORMAN, J. M., 1993, Scaling processes between leaf and canopy levels. In Scaling

    Physiological Processes. Leaf to globe, edited by J. R. Ehleringer and C. B. Field(New York: Academic Press, Inc.), pp. 4177.

    NORMAN, J. M., KUSTAS, W. P., and HUMES, K. S., 1995, Source approach for estimatingsoil and vegetation energy fluxes in observations of directional radiometric surface

    temperature. Agricultural and Forest Meteorology, 77, 263293.NOVAK, M. D., 1986, Theoretical values of daily atmospheric and soil thermal admittances.

    Boundary-Layer Meteorology, 24, 1734.PAW, U. K. T., and GAO, W., 1988, Applications of solutions to non-linear energy budget

    equations. Agricultural and Forest Meteorology, 43, 121145.QIU, G . Y. , YANO, T., and MOMII, K., 1998, An improved methodology to measure

    2550 E. Boegh and H. Soegaard

    y

    [

    g

    y

    ]

  • 8/2/2019 Remote Sensing Based Estimation of Evapotranspiration Rates 2004Boegh

    18/18

    evaporation from bare soil based on comparison of surface temperature with a drysoil surface. Journal of Hydrology, 210, 93105.

    SEGUIN, B., ASSAD, E., FRETEAUD, J. P., IMBERNON, J., KERR, Y., and LAGOUARDE, J. P.,1989, Use of meteorological satellites for water balance monitoring in Sahelianregions. International Journal of Remote Sensing, 10, 11011117.

    SELLERS, P. J., BERRY, J. A., COLLATZ, G. J., FIELD, C. B., and HALL, F. G., 1992, Canopyreflectance, photosynthesis and transpiration. III. A reanalysis using improved leafmodels and a new canopy integration scheme. Remote Sensing of Environment, 42,187216.

    SHUTTLEWORTH, W. J., and WALLACE, J. S., 1985, Evaporation from sparse cropsanenergy combination theory. Quarterly Journal of the Royal Meteorological Society III,469, 839857.

    SOEGAARD, H., JENSEN, N. O., BOEGH, E., HASAGER, C. B., SCHELDE, K., and THOMSEN, A.,2002, Carbon dioxide exchange over agricultural landscape using eddy correlationand footprint modeling. Agricultural and Forest Meteorology, 114, 153173.

    STEWART, J. B., KUSTAS, W. B., HUMES, K. S., NICHOLS, W. D., MORAN, M. S., andDE BRUIN, H. A. R., 1994, Sensible heat fluxradiometric surface temperaturerelationship for 8 semi-arid sites. Journal of Applied Meteorology, 33, 11101117.

    SUGITA, M., and BRUTSAERT, W., 1990, Regional surface fluxes from remotely sensed skintemperature and lower boundary layer measurements. Water Resources Research, 26,29372944.

    THORGEIRSSON, H., and SOEGAARD, H., 1999, Simulated carbon dioxide of leaves of barleyscaled to the canopy and compared to measured fluxes. Agricultural and ForestMeteorology, 98, 479489.

    TROUFLEAU, D., LHOMME, J. P., MONTENY, B., and VIDAL, A., 1997, Sensible heat flux andradiometric temperature over sparse Sahelian vegetation. I. An experimental analysisof the kB21 parameter. Journal of Hydrology, 188189, 815838.

    VAN DER KEUR, P., HANSEN, S., SCHELDE, K., and THOMSEN, A., 2001, Modification ofDAISY SVAT model for use of remotely sensed data. Agricultural and ForestMeteorology, 106, 215231.

    VERHOEF, A., DE BRUIN, H. A. R., and vAN DER HURK, B. J. J. M., 1997, Some practicalnotes on the parameter kB21 for sparse vegetation. Journal of Applied Meteorology,36, 560573.

    WANG, Y.-P., and LEUNING, R., 1998, A two-leaf model for canopy conductance,photosynthesis and partitioning of available energy I: Model description andcomparison with a multi-layered model. Agricultural and Forest Meteorology, 91,89111.

    YAMANAKA, T., and YONETANI, T., 1999, Dynamics of the evaporation zone in dry sandysoils. Journal of Hydrology, 217, 135148.

    YAMANAKA, T., TAKEDA, A., and SUGITA, F., 1997, A modified surface-resistance approachfor representing bare-soil evaporation: wind tunnel experiments under variousatmospheric conditions, Water Resources Research, 33, 21172128.

    Estimation of evapotranspiration rates 2551

    y

    [

    g

    y

    ]