Lim Etal2012

Embed Size (px)

Citation preview

  • 7/31/2019 Lim Etal2012

    1/13

    Agricultural and Forest Meteorology 152 (2012) 3143

    Contents lists available at SciVerse ScienceDirect

    Agricultural and Forest Meteorology

    journal homepage: www.elsevier .com/ locate /agr formet

    The aerodynamics ofpan evaporation

    Wee Ho Lim a , Michael L. Roderick a,b, , Michael T. Hobbins a,c , Suan Chin Wong a , Peter J. Groenevelda ,Fubao Sun a, Graham D. Farquhar a

    a Research School of Biology, The Australian NationalUniversity, Canberra, ACT 0200,Australiab Research School of Earth Sciences, The AustralianNational University, Canberra, ACT 0200,Australiac Colorado Basin River Forecast Center, NationalWeather Service, National Oceanic andAtmospheric Administration,Salt Lake City, UT 84116, USA

    a r t i c l e i n f o

    Article history:Received 12 April 2011

    Received in revised form 10 August 2011

    Accepted 19 August 2011

    Keywords:

    Pan evaporation

    Vapour transfer

    Boundary layer theory

    Aerodynamics

    a b s t r a c t

    In response to worldwide observations reporting a decline in pan evaporation over the last 3050 years,we developed an instrumented US Class A pan that replicated an operational pan at Canberra Airport in

    Australia. The aim of the experimental facility was to investigate the physics ofpan evaporation under

    non-steady state conditions. By monitoring the water level at 5-min intervals we were able to calculate

    the evaporation rate and thereby determine the short-term mass balance of the pan. Over the same

    time intervals, we also monitored (short- and long-wave) radiation, temperature (air, water surface, bulk

    water, inner and outer pan wall), atmospheric pressure as well as the air vapour pressure and the wind

    speed at a standard reference height (2 m above ground level). The experimental pan was operated for

    three years (20072010).

    In this paper, we develop a framework for quantifying vapour transfer by coupling Ficks First Law of

    Diffusion with boundary layer theory. This approach adequately represented pan evaporation measure-

    ments over short time intervals (half-hourly) under non-steady state conditions provided that surface

    temperature measurements, that account for the substantial cooling associated with evaporation, are

    available. It involved estimating the boundary layer thickness and other properties ofair above the evap-

    orating surface for a pan. Our results are consistent with the envelope oftheoretical curves concept for

    the wind function introduced by Thom et al. (1981).

    2011 Elsevier B.V. All rights reserved.

    1. Introduction

    Pan evaporation is the evaporation from a standard water-filled

    dish and is the most widely used physical measure of the evapora-

    tive demand of the atmosphere. Pan evaporation measurements

    have been widely used in agricultural meteorology due to their

    simplicity, low cost and proven ease of application for irrigation

    scheduling (Stanhill, 2002). Due to the widespread applications,

    evaporation pans in various forms have been deployed in many

    regions for at least the last several decades (Brutsaert, 1982). Anal-

    ysis of worldwide pan evaporation data has found changes, mostly

    declines (Roderick et al., 2009a,b) despite the trend of rising globalaverage airtemperature. This has becomeknown as the panevapo-

    ration paradox (Peterson et al., 1995; Brutsaertand Parlange, 1998;

    Roderick and Farquhar, 2002), because it has occured concurrently

    with the global warming. Reduction in irradiance (Roderick and

    Farquhar, 2002) and wind speed appear to have been major causes

    of this phenomenon (Roderick et al., 2007).

    Corresponding author at: Research School of Biology, The Australian NationalUniversity, Canberra, ACT 0200, Australia.

    E-mail address:[email protected](M.L. Roderick).

    For pans located above the ground, the surface area for heat

    transfer is larger than for mass transfer (Kohler et al., 1955; Riley,

    1966). Consequently, moststudies haveemployeda pan coefficient,

    typically 0.7 for a US Class A pan (Stanhill, 1976), by which panevaporation is multiplied to givea valuemore representativeof nat-

    uralevaporation, to account for this largely radiativeeffect (Linacre,

    1994). In terms of theaerodynamics of panevaporation, Thomet al.

    (1981) examined the roles of free and forced convection. Rotstayn

    et al. (2006) subsequently developed the PenPan model of pan

    evaporation by combining the radiative model ofLinacre (1994)

    with the aerodynamic model ofThom et al. (1981). The PenPan

    model has been used to attribute the cause of changes in pan evap-oration in both observations (Roderick et al., 2007; Shuttleworth

    et al., 2009) andin climate models(Johnson and Sharma, 2010). The

    derivation of the PenPan model assumed steady state conditions

    that require a typical integration period of around 1 week in sum-

    mer and upto 1 month inwinter (Roderick et al., 2009a). However,

    the radiative forumulation (Linacre, 1994) that underlies the Pen-

    Pan model, whilst physically based, has never been experimentally

    tested. Further, the aerodynamic formulation (Thom et al., 1981)

    has not, to our knowledge, been subject to an independent exper-

    imental test. These experimental tests have a high priority given

    the prominent role attributed to declines in windspeed (stilling)

    0168-1923/$ see front matter 2011 Elsevier B.V. All rights reserved.

    doi:10.1016/j.agrformet.2011.08.006

    http://localhost/var/www/apps/conversion/tmp/scratch_4/dx.doi.org/10.1016/j.agrformet.2011.08.006http://localhost/var/www/apps/conversion/tmp/scratch_4/dx.doi.org/10.1016/j.agrformet.2011.08.006http://www.sciencedirect.com/science/journal/01681923http://www.elsevier.com/locate/agrformetmailto:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_4/dx.doi.org/10.1016/j.agrformet.2011.08.006http://localhost/var/www/apps/conversion/tmp/scratch_4/dx.doi.org/10.1016/j.agrformet.2011.08.006mailto:[email protected]://www.elsevier.com/locate/agrformethttp://www.sciencedirect.com/science/journal/01681923http://localhost/var/www/apps/conversion/tmp/scratch_4/dx.doi.org/10.1016/j.agrformet.2011.08.006
  • 7/31/2019 Lim Etal2012

    2/13

    32 W.H. Limet al. / Agricultural andForestMeteorology 152 (2012) 3143

    and/or solar radiation in explanations of the world-wide decline in

    pan evaporation (Roderick et al., 2007, 2009b).

    Detailed physical investigations on pan evaporation have been

    rare. Notably,Jacobs et al. (1998) have studied the sub-daily ther-

    mal behaviour of a standard US Class A pan. They found that the

    water in the pan was generally well-mixed. Their pan was located

    on the ground and therefore in strong thermal contact with the

    soil surface. We expect that the finding of a well-mixed water

    body would also hold for a pan located on an elevated (

    150mm)

    wooden platform such as used in standard US Class A pan installa-

    tions. The question of any depression in surface temperature due

    to evaporation from a thin (1 mm or so) layer immediately belowthe surface (Ward and Stanga, 2001; Hisatake et al., 1993, 1995)

    was not explicitly addressed because Jacobs et al. (1998) used a

    Penman-style formulation for the energy balance where the sur-

    face temperature was eliminated from the underlying equations.

    More recently, the finding of well-mixed water within the pan has

    also been reported based on experiments using an instrumentedUS

    Class A pan (Martinez et al., 2006). In that research, the water sur-

    face temperature was assumed to be uniform over a40mm deeplayer. As noted above, this is unlikely to be true for an evaporat-

    ing surface (Ward and Stanga, 2001; Hisatake et al., 1993, 1995),

    but again, the Penman-style formulation used in that research

    (Martinez et al.,2006) effectivelyeliminated the watersurface tem-

    perature from the equations. A second caveat applicable to that

    study was that the external walls of the pan were insulated to min-

    imise heat transfer and thereby simplify the analysis. Therefore,

    that installation did not mimic standard evaporation pans.

    To contribute to a better understanding of the world-wide

    trends in pan evaporation, we have constructed an instrumented

    experimental pan that replicates existing standard installations.

    We installed specialised sensors onto a standard US Class A pan

    as used by the Australian Bureau of Meteorology (BoM). The water

    level, (short- andlong-wave)radiation, temperature (air,water sur-

    face, bulk water, inner and outer pan wall), atmospheric pressure

    as well as the air vapour pressure and the wind speedat a standard

    reference height (2m above ground level) were all monitored at

    5-min intervals for a three-year period (20072010).This paper describes the first step in the development of a new

    parameterisation for a Penman-type combination equation for pan

    evaporation. Here, we focus explicitly on the aerodynamic compo-

    nent of pan evaporation. To do that, we develop a framework for

    quantifying vapour transfer by coupling Ficks First Law of Diffu-

    sionwith boundarylayer theory assuming thatsurfacetemperature

    measurements are available.We investigatethe underlyingphysics

    of mass transfer and test our theory using data collected at the

    experimental pan. We relate our research to the ideas put forward

    by Thom et al. (1981).

    2. Theory

    Section 2.1 summarises existing vapour transfer equations

    employed in environmental applications. Section 2.2 derives a

    vapour transfer equation based on Ficks First Law of Diffusion.

    Section 2.3 describes the significance of boundary layer theory in

    vapour transfer. Section 2.4 links the boundary layer theory with

    Ficks First Law of Diffusion for vapour transfer and presents an

    approach to quantifying vapour transfer from an evaporation pan.

    2.1. Existing vapour transfer equations

    Daltons Law assumes that vapour moves from high to low par-

    tial pressure and is generally expressed as

    E(es

    (Ts

    ) ea

    (Ta

    ))=f

    v(es

    (Ts

    ) ea

    (Ta

    )) (1)

    where E [m s1] is the evaporation rate of liquid water in tradi-tional hydrologic units of depth per unit time, es(Ts) [Pa] is the

    vapour pressure at the evaporating surface, ea(Ta) [Pa] is the air

    vapour pressure at the same height that air temperature is mea-

    sured at andfv [m s1 Pa1] is the aerodynamic function. (Note thates(Ts) is taken to be the saturated vapour pressure at the surface

    temperature.)fv depends on both wind speed and the temperaturedifference between the evaporating surface and the air, especially

    whenwindspeeds are low (Thom et al., 1981). For simplicity(Thomet al., 1981), fv has usually been taken as a function of wind speedand Eq. (1) becomes

    E= f(u)(es(Ts) ea(Ta)) (2)

    where f(u) is known as the wind function. In terms of the

    widely used resistance terminology (Monteith,1965; Monteith and

    Unsworth, 2008), Eq. (1) can be rewritten as

    E= MwMa

    awPa

    (es(Ts) ea(Ta))ra

    0.622 awPa

    (es(Ts) ea(Ta))ra

    (3)

    where ra [s m1] is the aerodynamic resistance, Ma [kgmol1] is

    the molecular mass of air, Mw [kgmol1] is the molecular mass ofwater, a [kgm3] is the density of air, w [kgm3] is the densityof liquid water and Pa [Pa] is the atmospheric pressure. In practice,

    this is more or less equivalent to Eq. (2) since ra is mostly driven by

    the wind (Chu et al., 2010). In subsequent sections we examine the

    functional form of the relation between ra and wind speed.

    2.2. Ficks First Law of Diffusion

    Ignoring complications of inhomogeneity, a one-dimensional

    form of Ficks First Law of Diffusion (see Fick (1995) for a trans-

    lation) can be written as

    J= DdCdz

    (4)

    whereJ[molm2 s1] is the flux density,D [m2 s1] is the diffusioncoefficient, C[molm3] is the molar concentration andz[m] is thedistance. (Note: the negative sign implies that J is positive when

    diffusion is toward the lower concentration.) For vapour transfer,

    Eq. (4) can be expressed in a finite form assuming an ideal gas

    J= DvR

    (ea(Ta)/Ta) (es(Ts)/Ts)z

    = DvR

    (es(Ts)/Ts) (ea(Ta)/Ta)z

    (5)

    where Dv [m2 s1] is the diffusion coefficient for water vapour inair, R [J mol1 K1] is the ideal gas constant, Ts [K] is the water sur-face temperature, Ta [K] is the air temperature and z [m] is theboundary layer thickness (see details in Sections 2.3 and 2.4). For

    typical environmental conditions,

    es(Ts)

    Ts ea(Ta)

    Ta es(Ts) ea(Ta)

    Ta,

    and Eq. (5) can be simplified as

    J 1R

    DvTa

    (es(Ts) ea(Ta))z

    (6)

    This is Ficks First Law of Diffusion in Daltons form for vapour

    transfer (see Appendix D for relation to formulations commonly

    used in plant physiology). It should be noted that Dv is not a con-stant, but increases with Ta (Gilliland, 1934) and varies inversely

    withPa (e.g., Monteith and Unsworth, 2008; Rohsenow et al., 1985).

    An equation to calculate Dv as a function ofTa and Pa is given in

  • 7/31/2019 Lim Etal2012

    3/13

    W.H. Limet al. / Agricultural andForestMeteorology 152 (2012) 3143 33

    Fig.1. A schematic diagram of thethreshold model adopted herefor the variationin vapour pressure (e) with height (z) above the evaporating surface.

    After Leighly (1937) and Machin (1964, 1970).

    Appendix A. We can express Eq. (6) in units of depth per unit time

    (i.e., E) by incorporating Mw and w ,

    E= MwRw

    DvTa

    (es(Ts) ea(Ta))z

    (7)

    This is thegeneral form of FicksFirstLaw of Diffusionfor vapour

    transfer based on an ideal gas. By comparing Eqs. (1) and (7) we see

    that the aerodynamic function (fv in Eq. (1)) is

    fv =

    Mw

    Rw

    Dv

    Ta

    1

    z(8)

    and the conductancegm,v [mol m2 s1] is given by

    gm,v=1

    R

    DvPaTa

    1

    z(9)

    2.3. Boundary layer theory

    Thickness of the boundary layer in air has been measured

    directly at various times and by various methods. Its order of

    magnitude is from a few millimeters down to some small frac-

    tion of a millimeter. Within it, gradients of vapor concentration,

    temperature, and velocity are linear. Leighly (1937)

    Current conceptions of vapour transfer are for a thin (bound-ary) layer of air above the evaporating surface with transport of

    vapour across that layer by molecular diffusion (e.g., Giblett, 1921,

    pp. 473474; Penman, 1948) and this is confirmed by observa-

    tion (Doe, 1967). Further, the boundary layer thickness is known

    to decrease as the wind speed increases (Machin, 1964, 1970). The

    treatment consistent with those experimental results holds that

    the vapour pressure at the top of the boundary layer is the same as

    the vapour pressure at a reference height (e.g., 2m above ground

    level) (Leighly, 1937). The simplified threshold model is depicted

    in Fig. 1.

    To ensure that the framework adopted here can be applied at

    other evaporation pans we make the initial assumption that mea-

    surements of vapour pressure, air temperature and wind speed

    will be available at the reference height. Hence, the challenge is

    to estimate the boundary layer thickness zusing those availablemeasurements.

    2.4. Boundary layer thickness

    Vapour transfer can be conceived as due to free or forced con-

    vection, or a mixture of both, often called mixed convection

    (Monteith and Unsworth, 2008). It is difficult to specify precise

    boundaries between these various convection regimes. Here, wepropose a convenient structure for estimating zwithout a pri-

    ori selection of the thresholds. Following boundary layer theory

    (Hisatake et al., 1993,1995) we formulatethe boundary layerthick-

    ness as

    z=f(Re,L) ReqL (10)

    where Re is the Reynolds number (dimensionless) and L [m] is the

    characteristic length of the evaporating surface. For a cylindrical

    evaporation pan, we assume that L is the diameter (1.21m for a

    US Class A pan). Here q is a dimensionless constant (range: 0 to

    1.0). Conventionally, q is 0.5 for laminar flow over a flat surface(Schlichting, 1960).

    Traditionally, Re is calculated using the free stream velocity

    of the air and adjusted using a numerical factor (e.g., Eq. 2.2

    in Schlichting (1960)). Conceptually, the numerical factor is an

    attempt to estimate the wind velocity immediately adjacent to

    the evaporating surface. Importantly, the numerical factor ofRe

    will depend on the height at which wind velocity is measured. An

    alternative formulation is to calculate Re using the wind velocity

    immediately adjacent to the evaporating surface. To develop such

    an expression forRe we note that

    Re inertial forcesviscous forces

    = ausLa

    (11)

    where us [m s1] is a three-dimensional wind velocity immedi-

    ately adjacent to the evaporating surface anda [kgm1 s1] is thedynamic viscosity of air. (Note that a is a function of air tem-perature, see Appendix B.) Here, us

    f(uV, uH) where uV [m s

    1]and uH [m s

    1] are the vertical (analogous to free convection) andhorizontal (analogous to forced convection) components respec-

    tively. When uVuH, us is dominated by the vertical component,i.e., free convection dominates. Alternatively, when uHuV, us isdominated by the horizontal component, i.e., forced convection

    dominates. When uV and uH are of similar magnitude, mixed con-

    vection occurs.

    In our experiment, the wind velocity components above the

    evaporating surface (i.e.,uVanduH) arenot measured directly.Here

    we assume that uH is some fraction of the horizontal wind speed at

    the reference height, urefas follows

    uH= nuref (12)

    wheren is a dimensionless constant(range: 0 to1.0)anduref[m s1]

    is the horizontal wind speed measured at the reference height, e.g.,

    2 m above groundlevel. Using thestandard theorybased on thever-

    tical gradient in air density (see Appendix C for details), we derive

    that uV can be calculated as

    uV = kuV,C (13)

    where k is another dimensionless constant (range: 0) and uV,C[m s1] is the characteristic speed of air movement in the verticaldirection.

    One way of combining uH and uV to estimate us is

    us =uV

    + uH1/

    (14)

    where is a dimensionless constant (range: 1 to ). For exam-

    ple, is equivalent to the assumption that us is the maximum

  • 7/31/2019 Lim Etal2012

    4/13

    34 W.H. Limet al. / Agricultural andForestMeteorology 152 (2012) 3143

    of the free or forced convection component (McAdams, 1954; Ball

    et al., 1988). Alternatively, setting = 1 implies that us is the sum

    of free and forced convection components (Adams et al., 1990).

    Intermediate values between these extremes change the relative

    contributions ofuVand uH to us.

    Combining Eqs. (10)(14) the boundary layer thickness is given

    by

    z= a[(kuV,C) + (nuref)]1/

    L

    aq

    L (15)

    Substituting Eq. (15) into Eq. (7), the final form of the evapora-

    tion equation is

    E= MwRw

    DvTa

    (es(Ts) ea(Ta))

    [(a[(kuV,C) + (nuref)]

    1/L)/a]

    qL

    (16)

    3. Materials and methods

    3.1. Field installation

    The experimental US Class A Pan (with bird guard) was located

    at the BoM field station at Canberra Airport (Australia, Latitude:

    35.3S, Longitude: 149.2E, elevation 578m) and was directlyadjacent (5 m) to a BoM operational US Class A pan. Once oper-ational, our experimental pan was replenished with water daily

    (at 9am local time) by the BoM duty officer following standard

    operating procedure. We commenced installation in September

    2006 and the pan was operational from early 2007 until 20 January

    2010.

    The experimental pan facility is depicted in Fig. 2. The pan

    was equipped with a water-level sensor (see details below) that

    enabled us to calculatethe mass balance at 5-min intervals. In addi-

    tion, Pt100 temperature sensors (Type GW2105, dimension 2 mm

    2.3 mm, Degussa, Hanau, Germany) were located on the interiorand exterior pan wall at the four cardinal compass points, and at

    three different levels (25, 100 and 175mm from the bottom), to

    characterise the thermal dynamics of the pan. Temperature sen-sors were also placed at the same three levels at the centre of

    the pan to record the bulk water temperature. The temperature

    of the water surface was measured using an infrared thermome-

    ter (Model: M50-1C-06-L, Mikron Instrument Co. Inc., Oakland, NJ,

    USA) (Fig. 2b).

    The evaporation rate was calculated from water height mea-

    surements made using a magnetostrictive linear displacement

    transducer (MLDT) (MagneRule Plus, MRU-4001-015, Schawitz

    Sensors, Hampton, VA, USA) with a spherical float. The float was

    installed in a stilling well connected to one side of the pan. After

    initial experimentation, and contrary to the manufacturers spec-

    ifications, we found that the output of the MLDT was sensitive to

    variationsin ambient temperature. To overcome thatlimitation,we

    attached a proportional-integral-derivative (PID) controlled heaterto the casing of the sensor head to maintain a constant tempera-

    ture of 40 C. The resolution of the MLDT for our installation was10m.

    In addition, standard meteorological measurements included

    radiation, wind speed, air temperature, air vapour pressure and

    atmospheric pressure. All components of the radiation balance

    (incoming andoutgoingshort- and long-waveradiation) weremea-

    sured using a Kipp & Zonen CNR 1 Net Radiometer attached to

    a swinging-arm (Fig. 2). Most of the time, the swinging arm was

    parked to the southeast of the pan with the downward sensors fac-

    ing the ground. At 5-min intervals, a motor swung the arm over the

    centre of the pan where the downward sensors sampled radiation

    from the water surface for a 20-s period. Forty radiometer readings

    aretakenin each directional swing.(A paper focusing on theenergy

    balance ofpan evaporationis in preparation.) Wind speedwas mea-

    sured using a cup anemometer at 2 m above ground level (u2). In

    addition,a mast wasinstalled 5 m away from the experimentalpan

    to enable installation of temperature, vapour pressure and atmo-

    spheric pressure sensors and a 2D ultrasonic anemometer (Wind

    Observer II, Gill Instrument Ltd.). The 2D ultrasonic anemome-

    ter was located at the same height as the cup anemometer to

    enable us to check the performance of the cup anemometer. Atmo-

    spheric pressure was measured with a Vaisala Pressure Transmitter

    (Model: PTB101B). Air temperature and air vapour pressure were

    measured with a Vaisala Humitter (Type 50Y, Vaisala, Helsinki,

    Finland).Air temperature,vapour pressure andthe water level sen-

    sor were all calibrated in the laboratory and periodically checked

    on-site after field installation. All analog sensors signals were con-

    vertedvia a 16-bit analog-to-digital converter,averagedover 5-min

    intervals and stored in a single-board computer.

    3.2. Data sampling

    Short-term oscillations in the water level were apparent in the

    5-min data. To avoid the high-frequency noise, all water level data

    were aggregated, and the resulting evaporation rate calculated, at

    half-hourlyintervals. Thechange in water level dueto dailyrefilling

    (at9amlocaltime)was takeninto account. All othermeteorologicaldata were resampled to half-hourly intervals.

    From the database, we identified 160 days of elite data that

    had no missing half-hourly totals and included 40 (rainless) days

    each in spring, summer, autumn, andwinter, respectively. The final

    elite database contains (= 160 days 48 samples per day) a totalof 7680 half-hourly measurements.

    3.3. Parameter estimation and validation

    We split the datafromthe 160 daysintotwo databases. The first

    subset (60 days, including 15 days in each of the four seasons) was

    used to estimate the model parameters. The second subset (100

    days, including 25 days in each of the four seasons) was used to

    validate the model.While theoptimumvalue of (Eq. (16)) was >2, theeffect on the

    fit to the data was not strong, and we chose=2 based ona vecto-

    rial combination assumption (Adams et al., 1990). The parameters

    (k, n, q) were estimated using 2880 half-hourly measurements (=

    60 days 48 samples per day) using a least squares optimisationapproach. To do that we initially assumed values of the parame-

    ters and then computed pan evaporation Epan for that parameter

    combination. Many possible combinations were tested using an

    automated computer algorithm and the parameter combination

    with the lowest root mean square error (RMSE) was selected.

    4. Results

    4.1. Meteorological data, calibration andvalidation

    The wind speed measured by the cup anemometer (u2, also

    known as the (horizontal) wind speed hereafter) was compared

    to measurements by the 2D ultrasonic anemometer. We used half-

    hourly data from all 160 days (7680 samples) (Fig. 3). At low wind

    speeds (

  • 7/31/2019 Lim Etal2012

    5/13

    W.H. Limet al. / Agricultural andForestMeteorology 152 (2012) 3143 35

    Fig. 2. Experimentalpan installationat Canberra Airport BoMstation. (a)The US Class A pan is 1.21m in diameter (4ft) and 0.254m in height (10 in.), equipped with instru-

    ments measuring (short- and long-wave) radiation, wind, temperature(water surface, bulk water, inner and outer panwall) and water level. (b) The infrared thermometer

    installation (inside thewhite PVC pipe facingthe watersurface in thepan) measuring long-waveradiation emitted by thewatersurface.

    Fig. 3. Comparison of wind speed measurements from the cup anemometer and the 2D ultrasonic anemometer over half-hourly intervals at the experimental pan (7680

    data points,y=1.07x0.44, R2 =0.97, RMSE= 0.42ms1).

    The least squares estimates of the parameters are k=0.20,

    n= 0 .10, and q=0.64 (half-hourly samples: 2880, regressionof estimated versus observed Epan: slope = 0.86, inter-

    cept=6.7106

    mm s1

    , R2

    =0.82, RMSE=3.1105

    mm s1

    ).

    The estimate for q is close to the previously noted value of0.5for laminar flow. The estimate ofn is also sensible in that it makes

    the horizontal component of wind velocity immediately above the

    surface only10% of the wind speed at2 m above ground level. With

  • 7/31/2019 Lim Etal2012

    6/13

    36 W.H. Limet al. / Agricultural andForestMeteorology 152 (2012) 3143

    Fig. 4. Pan evaporation Epan (observed versus estimated), conductancegm,v (observed versus estimated per Eq. (17)), vapour pressures (es(Ts), ea(Ta)), temperatures (Ts , Ta),

    u2 (wind speedat 2m aboveground level) and atmospheric pressure Pa for theexperimental pan fordiurnal cyclesin: spring (a andb), summer (c and d),autumn (e andf),

    and winter (g and h).

  • 7/31/2019 Lim Etal2012

    7/13

    W.H. Limet al. / Agricultural andForestMeteorology 152 (2012) 3143 37

    Fig. 4. (Continued ).

  • 7/31/2019 Lim Etal2012

    8/13

    38 W.H. Limet al. / Agricultural andForestMeteorology 152 (2012) 3143

    Fig. 5. Estimated versus observed Epan by integrating half-hourly results to a daily basis for 100 days: (a) Ts available (y=1.04x+0.39, R2 =0.98, RMSE=0.51mmd1); (b)

    assume Ts =Ta (y= 1.56x+1.43, R2 =0.91, RMSE=3.06mmd1).

    Fig. 6. Temperature depression of water surface at half-hourly intervals for the experimental pan (7680 data points). (a) Frequency distribution ofTs Ta (binsize= 0.1K).(b) Frequency distribution ofTs Tw (binsize= 0.1K). (c) Observed Epan versus Ts Ta .

    those results, the pan evaporation Epan (semi-empirical equation

    from Eq. (16)) is

    Epan =MwRw

    DvTa

    (es(Ts) ea(Ta))a

    (0.20uV,C)2 + (0.10u2)2L

    /a

    0.64L

    (17)

    Note that the numerical value ofL is 1.21m, i.e., the diameter of

    a US Class A pan.

    We subsequently used Eq. (17) to estimate Epan for the remain-

    ing (4800) half-hourly totals. (The resulting half-hourly totals are

    compared with measurements over eight typical days in Fig. 4.)

    The half-hourly totals were then summed into 100 daily totals and

    compared with measurements (Fig. 5a). The model explained 98%

    of the observed variance in the daily Epan with an overall RMSE of

    0.51mmd1 giving us confidence in the parameterisation.

    4.2. Water surface temperature and evaporative cooling

    The model parameters and results to date were derived using

    our (infrared) measurement of the water surface temperature Ts.

  • 7/31/2019 Lim Etal2012

    9/13

    W.H. Limet al. / Agricultural andForestMeteorology 152 (2012) 3143 39

    Fig.7. Half-hourlyboundary layerthickness z(perEq. (15);= 2,k= 0.20,n= 0.10,

    q=0.64) versus u2 (wind speed at 2m above ground level) for the experimentalpan (7680 data points).

    Fig. 8. Half-hourly aerodynamic functionfv (per Eq. (17)) versus u2 (wind speed at

    2 m aboveground level)for theexperimental pan (7680 data points).

    In practical applications, and especially for evaluating the histori-

    cal records, Ts is unknown. Further, inspection ofFig. 4 shows that

    temperature of the water surfaceTs can be quite different from thatof the air Ta, as would be expected for a freely evaporating surface.

    The following question arises: canEpan be estimatedaccuratelywith-

    out measurementsof Ts in theabsence of radiationmeasurements? To

    answer this, we assumed that the water surface was at the air tem-

    perature (i.e., Ts = Ta) and accordingly recalculated the saturated

    vapour pressure at the surface. The results, using totals from the

    100 elite-days, show that this is a bad assumption for a purely

    aerodynamic formulation of evaporation (Fig. 5b). In general, the

    estimateof dailyEpan basedontheassumptionthat Ts =Ta was much

    larger than the observations at high rates ofEpan. That result means

    that the water surface is cooler than the air when Epan is high and

    implies evaporative cooling. The same phenomenon is visible in

    Fig. 4, where Ts is substantially lower than Ta in the mid-afternoon

    when Epan tends to be highest.

    To investigate further, we examined the relationships between

    Ts, Ta and the bulk water temperature Tw (taken as the average of

    measurements at 25, 100 and 175mm below the water surface).

    The bulk water was found to be well mixed (results not shown) but

    the evaporating surface was often up to 3 K warmer or 5K cooler

    than the bulk water, with an overall average of around 2K cooler

    than the bulk water (Fig. 6b). We also found that the evaporating

    surface could be up to 5K warmer or 11K cooler than the air, and

    was on average, 1K cooler than the air over the 160 days (Fig. 6a).

    In general, the temperature of the water surface was much lower

    than the air when Epan was high (Fig. 6c), confirming our earlier

    deductions about evaporative cooling. In summary, the magnitude

    of the surface cooling due to evaporation was substantial.

    4.3. Estimates of boundary layer thickness and aerodynamic

    function

    Estimates of boundary layer thickness zusing the estimatedparameter values are plotted as a function of wind speed at 2m

    above ground level (u2) in Fig. 7. For u2> 1 m s1, the results are

    dominated by forced convection with z in the range 14 m m.For u2< 1 m s

    1, zis highlyvariablewithinan envelope constraintimposed by the free convection regime.

    The resulting aerodynamic function has been computed usingall available half-hourly data (Fig. 8). The overall features of the

    aerodynamic function are consistent with the ideas put forward by

    Thom et al. (1981, Figs. 2 and 5). In particular, at low wind speeds

    (u2 < 1 m s1) we see the variation in the aerodynamic function due

    to a mixture between free and forced convection. At higher wind

    speeds when forced convection dominates, the relation collapses

    tobe u20.64, which is nearly a straight line.

    4.4. Estimates of pan evaporation without a bird guard

    Our experimental pan used the same bird guard as used by the

    Australian BoM (Fig. 2). The effect is to reduce the mass transfer

    by

    7% in comparison to a similar pan without a bird guard (van

    Dijk, 1985). For comparative purposes, it is useful to estimate theimpact of the bird guard. On the boundary layer formulation used

    here (Fig. 1), the bird guard would affect the evaporation rate by

    reducing the wind speed near the evaporating surface and thereby

    increasingthe boundarylayer thicknessz. Hence,we canincorpo-rate that effectby adjusting thenumerical value of thenparameter.

    Assuming all else is held constant, we find a decrease in n of close

    to 10% will reduce Epan by around 7%. In summary, in the absence

    of a bird guard, the parameter estimates are k= 0.20, n= 0.11,

    q=0.64.To compare our adjusted formulation for a pan without a bird

    guard with previous research, we plotted the aerodynamic func-

    tion fv as a function of wind speed (2 m above ground level) u2for various differences between the water surface temperature Ts

    and the air temperatureTa under typical conditions (Pa =101.3kPa,ea(Ta)=1kPa, Ta = 293.15K (20

    C)). The resulting envelope of the-oretical curves (for differentTs Ta) (Fig. 9a) is consistent with theconcepts proposed by Thom et al. (1981, Figs. 2 and 5). This enve-

    lope coversthe majority(butnot all) of aerodynamic functionsused

    in previous studies when free or mixed convection dominates (i.e.,

    u2 < 1 m s1); and has a linear or near-linear form of the classical

    wind functions under high wind speeds when forced convection

    dominates (Fig. 9b).

    5. Discussion and summary

    The theoretical formulation derived here is based on the idea

    of vapour transport, predominantly by diffusion, from the liquid to

    vapour phases across a well-defined boundary layer (Fig. 1). This

  • 7/31/2019 Lim Etal2012

    10/13

    40 W.H. Limet al. / Agricultural andForestMeteorology 152 (2012) 3143

    Fig. 9. Aerodynamic functionfv versus u2 (wind speed at 2m above ground level) for pan evaporation under typical conditions (Pa = 101.3 kPa, ea(Ta)=1kPa, Ta =293.15K

    (20 C)). (a) Our model without a bird guard (per Eq. (16); = 2, k=0.20, n=0.11, q=0.64) for various differences between water surface and air temperature; (b) fv fromprevious studies.

    is well established in laboratory studies (Doe, 1967; Machin, 1964,

    1970) and the resulting formulation (Eq. (7) has a direct physical

    interpretation. Hence, the utility of this approach rests on whether

    the conceptual framework (Fig. 1) is useful.

    The assumption that= 2 is basedon theideaof vector average

    of the fluxes associated with the two physical processes (Adams

    et al., 1990). In contrast, the idea of taking a maximum value of

    free or forced convection (McAdams, 1954; Ball et al., 1988) means

    having , which might be useful over very short time inter-vals (e.g., minutes); yet could be less appropriate over longer time

    intervals (e.g., half hours) since a mixture of both free and forced

    convection is most likely to be the case under outdoor conditions.

    Typical wind function approaches of the formf(u) =a+bu implic-

    itly assume the free convection component (a) to be a constant(Adams et al., 1990). This is similar to setting = 1 and q=1 inEq. (16).

    Once was set = 2, the remaining parameters were estimated

    using the measurement database consisting of half-hourly data for

    60 days. The resulting parameters(k=0.20, n=0.10, q=0.64) werecloseto broad expectations. The estimatedkuV,Crange(00.6 m s

    1)is within our expected range ( 1 m s1), forced convection dominates, and

    the air above the pan is quickly replaced by the surrounding air.

    Under those conditions, z is predominantly a function of wind

    speed. However, under low wind speeds (u2 < 1 m s1), free con-

    vection dominates, and other factors also play important roles in

    determiningz. In particular, the spatial gradientof temperature inthe vertical direction (a surrogate for density difference) becomes

    important at low wind speeds.

    Our theory (Eq. (17)) and subsequent results (Fig. 8) show that

    a unique wind function does not exist. It also suggests a depen-

    dence on atmospheric pressure.Thus it would be interesting to test

    whether the formulation presented here (Fig. 9a) could be applica-

    ble under a wider set of conditions, such as at high altitude sites

    (Blaney, 1960; Giambelluca and Nullet, 1992) where the atmo-spheric pressure is substantially reduced. Our formulation assumes

    that the water surface is always close to the top of the pan and

    thereby avoids the shelter effect (Chu et al., 2010). This is most

    easily achieved in operational settings by refilling the pan each

    day.

    Measuring the water surface temperature Ts (in absence of

    radiation measurements) proved to be important for accurately

    estimating Epan using the aerodynamic approach when the evapo-

    ration rate was high and the associated evaporative cooling of the

    surface was at a maximum (Figs. 5 and 6). On average, the evap-

    orating water surface was cooler than both the air and the bulk

    water (Fig.6a and b), as foundpreviously in laboratory experiments

    (Hisatake et al., 1995, Fig. 5). This implies that the very thin layer of

    surface water, is, on average, a net absorber of sensible heat fromboth the air above it and from the bulk water below it. We were

    surprised by the magnitude of this cooling effect.

    Acknowledgements

    We thank the BoM staff; Tony McCarthy, Ross Hearfield, Kirsty

    Rhind, Nigel Smedley, David Pottage, Neil McArthur and Kenn Batt

    for their help in maintaining our experimental pan at the Canberra

    Airport and Liang Li for his contribution in setting up the exper-

    imental pan database. We acknowledge the Australian Research

    Council (ARC) for the financial support of this study through the

    grant DP0879763. We are grateful to two anonymous reviewers

    for helpful comments.

  • 7/31/2019 Lim Etal2012

    11/13

    W.H. Limet al. / Agricultural andForestMeteorology 152 (2012) 3143 41

    Fig. A.1. Diffusion coefficient Dv versus air temperature Ta at different values of

    atmospheric pressurePa.

    Appendix A. Diffusion coefficient of water vapour

    The diffusion coefficient of water vapourDv is calculated basedon Pruppacher and Klett (1997):

    Dv = 2.11

    Ta273.15

    1.94PoPa

    105 [m2 s1] (A.1)

    where Ta [K]is the air temperature,Pa [Pa] is the atmospheric pres-

    sure, Po [Pa] is the atmospheric pressure at the mean sea level

    (101.325 kPa). Fig. A.1 shows the change ofDv with Ta and Pa using

    Eq. (A.1).

    Appendix B. Dynamic viscosity of air

    The dynamic viscosity of air a (assumed dry air for simplicity)is calculated based onJacobson (2005):

    a = 1.8325

    416.16

    Ta + 120

    Ta

    296.16

    1.5 105 [kgm1 s1] (B.1)

    Although a is based on dry air instead of moist air, the differ-ence is small (Maxwell, 1866; Kestin and Whitelaw, 1964). Fig. B.1

    illustrates the change ofa with Ta using Eq. (B.1).

    Appendix C. Derivation of the speed of air in the vertical

    direction above evaporating surface

    The vertical circulation of air above the evaporating sur-

    face is determined by the air density difference (Schlichting,1960; Incropera and DeWitt, 1990; Holman, 2002; Monteith

    and Unsworth, 2008), which results from temperature gradients,

    vapour concentrationgradients,or a combination of both(Monteith

    and Unsworth, 2008).

    In principle, the speed of air in the vertical direction can be

    derived from Reynolds and Grashof numbers using dimensional

    analysis. An equivalent Reynolds number in the vertical direction

    (ReV) is the ratio of inertial forces (in the vertical direction) to vis-

    cous forces, i.e.,

    ReV =auVL

    a(C.1)

    where a [kgm3] is the density of air, uV [m s1] is the speed of

    air in the vertical direction, L [m] is the characteristic length of the

    Fig. B.1. Dynamic viscosity of aira versus air temperature Ta.

    evaporating surface anda [kgm1 s1] is the dynamic viscosity ofair (calculation ofa is given in Appendix B).

    The Grashof number (Gr) is equal to buoyancy forces times iner-

    tia forces divided by the square of viscous forces. Since the origin

    ofGr is in heat transfer studies (Karwe and Deo, 2003), it is com-

    monly calculated based on a spatial temperature difference. Here,

    we calculateGrby using thespatial density difference. The purpose

    is to take into account both temperature and concentration (water

    vapour and dry air) differences between the evaporating surface

    and the reference height, i.e.,

    Gr= 2ag(a/a)L

    3

    2

    a

    (C.2)

    whereg[m s2] is the gravitational acceleration, a [kgm3] is theaverage air density between the evaporating surface and the refer-

    ence heightanda [kgm3] is thespatialdifference in thedensityof air between the evaporating surface and the reference height.

    Assuming constant atmospheric pressure between the evapo-

    rating surface and the reference height, uV [m s1] can be derived

    from Eqs. (C.1) and (C.2) as follows,

    Re2V Gr, ReV Gr,

    auVL

    a

    2ag(a/a)L3

    2a,

    uV gaa

    L, uV = kgaa

    L, uV = kuV,C (C.3)

    where k is a dimensionless constant (range: 0) and uV,C [m s1]g(a/a)L

    is the characteristic speed of air in the verti-

    cal direction. We calculate a and a as

    a =1

    R

    (Pa ea(Ta))Ma + ea(Ta)Mw

    Ta

    (Pa es(Ts))Ma + es(Ts)MwTs (C.4)

  • 7/31/2019 Lim Etal2012

    12/13

  • 7/31/2019 Lim Etal2012

    13/13

    W.H. Limet al. / Agricultural andForestMeteorology 152 (2012) 3143 43

    Stanh ill, G. , 20 02. Is th e C lass A evapo ration pan still the most pr actical andaccurate meteorological method for determining irrigation water require-ments? Agric. For. Meteorol. 112, 233236, doi:10.1016/S0168-1923(02)00132-6.

    Stokes, V.J., Morecroft, M.D., Morison, J.I.L., 2006. Boundary layer conductance forcontrasting leaf shapes in a deciduous broadleaved forest canopy. Agric. For.Meteorol. 139, 4054, doi:10.1016/j.agrformet.2006.05.011.

    Thom, A.S.,Thony, J.L., Vauclin, M., 1981. On the proper employment of evaporationpans andatmometersin estimating potentialtranspiration.Quart.J. R.Meteorol.Soc. 107, 711736, doi:10.1002/qj.49710745316.

    van Dijk, M.H., 1985. Reduction in evaporation due to the bird screen used in theAustralian class A pan evaporation network. Aust. Meteorol. Mag. 33, 181183.

    Ward, C.A., Stanga, D., 2001. Interfacial conditions during evaporation or condensa-tion of water. Phys. Rev. E 64, 051509, doi:10.1103/PhysRevE.64.051509.