• Tidak ada hasil yang ditemukan

Directory UMM :Data Elmu:jurnal:A:Advances In Water Resources:Vol24.Issue2.2000:

N/A
N/A
Protected

Academic year: 2017

Membagikan "Directory UMM :Data Elmu:jurnal:A:Advances In Water Resources:Vol24.Issue2.2000:"

Copied!
11
0
0

Teks penuh

(1)

Directional radiometric temperature pro®les within a grass canopy

Russell J. Qualls

a,*

, David N. Yates

b a

Biological and Agricultural Engineering, University of Idaho, P.O. Box 440904, Moscow, ID 83844-0904, USA b

Civil, Environmental, and Architectural Engineering, University of Colorado at Boulder, Campus Box 428, Boulder, CO 80309-0428, USA

Received 4 January 1999; received in revised form 2 June 2000; accepted 18 June 2000

Abstract

The scalar roughness for sensible heat ¯ux,z0h, which appears in Monin and Obukhov similarity theory has been shown to exhibit substantial variability. Brutsaert and Sugita (cf. W. Brutsaert and M. Sugita, J. Atmos. Sci. 53 (1996) 209±16) derived an expression forz0hto account for variability inz0h as a result anisothermal vegetation skin temperature, which assumes an expo-nential pro®le of canopy skin temperature. We examined this assumption through directional radiometric measurements of tem-perature pro®les in a relatively uniform grass canopy. The temtem-perature gradient steepened monotonely throughout the morning and was modeled successfully with an exponential pro®le. The root-mean-square error (RMSE) between measured and modeled pro®les was small and ranged between 0.038°C and 0.530°C. One of the most signi®cant ®ndings of this research is that the decay coecient,

b, of the exponential pro®le model was approximately constant with time for this canopy. This may provide a useful simpli®cation in applying Brutsaert and Sugita's (loc. cit.) parameterization forz0h. Ó 2000 Elsevier Science Ltd. All rights reserved.

Keywords:Radiometric surface temperature; Remote sensing; Scalar roughness; Vegetation; Sensible heat ¯ux

1. Introduction

The exchange of mass and energy between land and atmosphere has signi®cant e€ects on the hydrology and weather of a region. These ¯uxes can be modeled with varying degrees of complexity with sets of energy bal-ance and continuity equations commonly referred to as land surface models (LSMs) or surface-vegetation± atmosphere transfer schemes (SVATs).

It is desirable to use directional radiometric surface

temperatures,hs, in these models because of the spatial

information that they contain. A directional radiometric temperature (DRT), is a radiometric surface tempera-ture measurement made along a speci®c sensor zenith view angle (i.e., the angle measured with respect to a vertically downward view). It is derived from a radiative energy balance of the surface using a radiance measurement from a single direction with a narrow ®eld of view, and provides the best approximation of the thermodynamic temperature based on a measurement of radiance [14]. However, vegetated surfaces impose

dif-®culties on the modeling process. In general, hs values

for a vegetated surface are not unique due to the com-plicated structure and the vertical and horizontal dis-tribution of the foliage skin temperatures. A marked vertical distribution of canopy surface temperature re-sults under conditions of strong shortwave radiation and moderately sparse leaf area density.

Complex multilayer canopy models have been de-veloped to accommodate this vertical complexity [13,20,26]. However, these complex models introduce a signi®cant computational burden, and stringent re-quirements in terms of initialization and updating that are undesirable especially for operational weather models, which cover subcontinental to continental scales.

Other models take a lumped approach and treat the surface as a single ``big leaf''. That is, they consider the canopy to be an isothermal entity [15,19]. Many of these models employ a bulk aerodynamic relationship for

surface sensible (H) heat ¯ux. An example is given by

Chen et al. [3] as

H ˆqcpChua…hsfcÿh† ˆqcp…hsfcÿh†=rh; …1†

where q is the air density; cp the coecient of speci®c

heat for air at constant pressure;Ch the bulk exchange

coecient for heat; ua the wind speed; hsfc the bulk

surface temperature;hthe temperature of the air; andrh

www.elsevier.com/locate/advwatres

*

Corresponding author. Tel.: 6184; fax: +1-208-885-7908.

E-mail address:rqualls@uidaho.edu (R.J. Qualls).

(2)

is the aerodynamic resistance to heat transfer,

rÿ1

h ˆChua.

Ch may be formulated directly from similarity theory

in terms of atmospheric stability correction functionswm

andwh for momentum and heat, respectively, and

sur-face roughness lengths. When distinct roughness lengths

z0mandz0hfor momentum and heat are retained,Chcan

be expressed as

wherez is the measurement height for wind speed and

air temperature, k the von Karman's constant, L the

Obukhov length, andR, estimated at 1.0, is the ratio of

the momentum and heat exchange coecients in the neutral limit.

The scalar roughness z0h which appears in (2) is an

important quantity. Verhoef et al. [27] notes that all meteorological weather models currently include a

parameterization for z0h, usually in the form of ln

…z0m=z0h†. Furthermore, recent tests of the Pan and

Mahrt [15] LSM by Chen et al. [3] show these bulk equations to be highly sensitive to the parameterization ofz0h.

Numerous parameterizations of z0h appear in the

literature. Garratt and Francey [6] proposed a constant

ratio z0m=z0hˆe2. Chen et al. [3] employed the

formu-lation of Zilitinkevich [29], z0m=z0hˆexpfkCpReg, in

the National Center for Environmental Prediction (NCEP/NOAA) Eta model, which is an operational

mesoscale model. HereCis an empirical constant; and

Re the roughness Reynolds number which describes

turbulence intensity. Using FIFE data, Qualls and

Hopson [18] employed z0m=z0hˆ expfb1‡b2…Re=

LAI†b‡b3Ug, where LAI is leaf area index, thebis and

b were empirically determined constants and U was a

function of daily maximum solar zenith angle. Although the constants were determined empirically, this rela-tionship is mathematically consistent with the

expres-sion forz0h derived by Brutsaert and Sugita [2]. It does

not account for diurnal variation of solar elevation nor does it account for sensor view angle e€ects. Stewart

et al. [22] analyzed values ofz0m=z0hobtained from eight

di€erent ®eld experiments. Kustas et al. [10] tested a

number of empirical expressions including z0m=z0hˆ

expfbua…hsÿh†g where hs was a DRT of the surface.

Other examples exist (e.g., Qualls and Brutsaert [17]:z0h

is a function of LAI; Sun and Mahrt [25]: z0h is a

function of …hsÿh†=H; Sugita and Brutsaert [23] and

Sugita and Kubota [24]: z0h is a function of solar

ele-vation; Matsushima and Kondo [13]: the aerodynamic

conductance, which is related to the bulk transfer

coef-®cient,Ch and z0h, is taken as a function of LAI, wind

speed and atmospheric stability).

Many of these studies show that z0h is neither

con-stant nor proportional to the momentum roughness length [3,10,13,17,18,25,27]. In order to accommodate

this variability ofz0h, Chen et al. [3] modi®ed the LSM

of Pan and Mahrt [15] by parameterizing z0h as a

function of roughness Reynolds number. The revised LSM has been implemented into the NCEP operational mesoscale Eta model.

Although the revised LSM simulated spatially aver-aged directional radiometric temperatures (DRTs) from FIFE reasonably well, there are instances where the simulated and measured DRTs di€ered by as much as 2±

C (see Fig. 5(d), panel C in Chen et al., [3]). When

di€erences occur in the comparisons provided, the model tends to under-predict mid-day DRTs, and

over-predict nighttime DRTs. Due to its sensitivity to z0h,

when the model is used in a mode in which DRTs serve as input to calculate sensible heat ¯uxes, these temper-ature di€erences would produce signi®cant errors in the modeled sensible heat ¯uxes. Studies by Qualls and Hopson [18] and Verhoef et al. [27] indicate that

parameterization of ln …z0m=z0h† by means of the

roughness Reynolds number alone leaves some unex-plained variability.

Some of this variability may be explained by vertical skin temperature gradients within the canopy and a distinct ground surface temperature. These cause the value of a DRT measured over a given surface to depend on the sensor view angle [8,12,28].

Fig. 1 illustrates the e€ect that a vertical foliage skin temperature gradient has on scalar roughness. Bulk formulations such as (3), shown later, operate under the assumption that the logarithmic air temperature pro®le

haextends down into the canopy (see single-dashed line

in Fig. 1), as if the atmospheric surface sublayer (SSL) extended all the way to the surface.

(3)

In reality, the canopy interferes with many of the processes so that we designate the region below the top of the vegetation as the canopy sublayer (CSL). No-tably, both the actual temperature pro®le of the air

within the canopy (ha(f)) and the foliage skin

tempera-ture (hf(f)) are disturbed from thehapro®le, wherefis a

non-dimensional depth into the canopy whose value is

zero atzˆh, and one atzˆ0. As a result, the ¯uxes at

depth f within the CSL are modulated by the

…hf…f† ÿha…f††di€erence.

When a DRT of the canopy is used for the surface temperature in the bulk formulation, the assumed

``height''z0h corresponding to this temperature must be

the height, where that value of hs would occur on the

extrapolated logarithmichapro®le. This may be seen by

the dash-double-dotted line extending down fromhs to

ha and then across to z0h in Fig. 1. As noted above, hs

depends on both the skin temperature pro®le through the canopy and on the view angle of the sensor, hence

z0h must be variable.

Under sparse canopy conditions, the view angle e€ect may be dominated by a combination of two ``sources'': a soil surface temperature and a single canopy tempera-ture. ``Dual-source'' models have been developed to address these conditions [5,11]. However, with more dense canopies, including some grasslands, vertical skin temperature variations through the canopy itself may have a signi®cant e€ect [13,20,21,26]. In order to address problems related to these vertical temperature gradients, canopy radiative transfer algorithms are often used. Based on the results of coupled canopy radiative

trans-fer and climate models, Prevot et al. [16] found z0h to

depend on sensor view angle. They concluded that their model produced reasonable results but that ®eld re-search was needed to verify the model. Furthermore, coupling a canopy radiative transfer algorithm to an LSM introduces a substantial computational burden.

Many experimental data sets include DRTs measured in situ, and from aircraft and satellite platforms. Some even have data for which DRTs were measured over the same point on the ground with sensors oriented at dif-ferent view angles [28]. However, these data sets require one to infer the nature of the canopy temperature pro-®les or to generate the propro-®les by modeling [4]; the pro®les cannot be validated by means of these data. This is not to say there have not been studies of the view angle e€ects on DRT [8,11±13,21,26], however, only a few studies exist, where canopy temperature pro®les were measured [9].

Despite studies that have explicitly addressed the ef-fects of view angle on DRT, there is no clear consensus on a method to overcome the problems associated with it. Some recent studies have shown that certain optimal view angles exist which minimize the view angle e€ects

[2,13]. These typically range between 50°and 70°from

nadir. However, these optimal angles are of little

prac-tical value since surface temperature data from satellite remote sensing images are measured across a range of

angles, usually less than 50°.

In order to receive the bene®t of spatial continuity which may be achieved through the use of spatially-distributed DRTs, while maintaining computational ef-®ciency of a single canopy layer in an LSM, one must accommodate the anisothermal canopy pro®le and sensor view angle e€ects. Recent theoretical advance-ments in embedding the CSL e€ects within a bulk transfer equation by Brutsaert and Sugita [2], which have been furthered by Qualls and Hopson [18] and Crago [4], have created the opportunity to account for these e€ects more rigorously than was previously possible. To aid these developments, experimental studies regarding the representation of the canopy temperature pro®le are crucial.

The purpose of this paper is to present and model ®eld observations of plant canopy DRT pro®les and to show some realistic values of temperature gradients and their corresponding semi-diurnal variability that might be expected. In addition, we will conduct a sensitivity analysis of Brutsaert and Sugita's [2] analytical

expres-sion for scalar roughnessz0h, particularly with respect to

realistic shapes and magnitudes of vertical skin tem-perature gradients within a plant canopy.

2. Methods

2.1. The anisothermal scalar roughness length

In practical applications, there are several possible ways to parameterize the heat transfer characteristics of a land surface covered with vegetation. One of the simpler methods is to use the bulk transfer approach given by Monin±Obukhov (M±O) similarity theory [1], which includes the scalar roughness length for sensible

heat, z0h, given as

where his the potential temperature of the air at some

reference height zwithin the surface sublayer (SSL); hs

an e€ective or bulk surface temperature intended here to

be supplied by DRT;aÿ1

h the turbulent Prandtl number

(which is a constant close to unity); k von Karman's

constant;uthe friction velocity;qthe air density;cpthe

coecient of speci®c heat at constant pressure; d0 the

zero plane displacement height; Lthe Obukhov length;

and wh the stability correction function for the

tem-perature pro®le in the SSL.

Unfortunately, z0h is neither constant, nor

propor-tional to the momentum roughness length,z0m as noted

earlier. In order to derive an expression for z0h for

(4)

and solved a di€erential equation for within-canopy scalar transport, and matched it to the M±O similarity ¯ux pro®le equation for the SSL. The main results of their work were expressions for the isothermal and

an-isothermal roughness lengths,z0hiandz0h, respectively,

z0hˆz0hiexp

where z0m is the momentum roughness; h the canopy

height; C2ˆ …2LAICtfh†=…ahk…hÿd0††; LAI the leaf

area index; Ctf the bulk transfer coecient for foliage

elements introduced in Brutsaert and Sugita [2];

r2ˆ ‰aÿ …a2‡4C2†1=2Š=2;a the extinction coecient in

the exponential shear stress pro®le and in the eddy dif-fusivity pro®le, which are assumed equal and have val-ues between 2 and 4 for dense ¯exible elements (wheat,

oats, immature corn);bis the temperature pro®le decay

coecient described below;Hthe sensible heat ¯ux; and

wa temperature-weighting coecient which relates the

view angle dependent DRT to the foliage temperature at the top and bottom of the canopy, as discussed below.

Note that z0hi only depends on the bulk canopy

char-acteristics.

In their derivation, Brutsaert and Sugita [2] assume a vertically uniform canopy structure for which the foliage skin temperature in the canopy can be represented with an exponential decay model

hf…f† ˆhfg‡ …hfhÿhfg†eÿbf; …6†

where hf(f) is the foliage skin temperature at depth f

within the canopy de®ned asfˆ1ÿz=hwherezis the

height above the ground surface, and h is the canopy

height.fvaries from 0 at the top of the canopy to 1 at

the ground.hfhis the foliage skin temperature at the top

of the canopy andhfg is the asymptotic limit of the

fo-liage temperature far below the bottom of the canopy as

fbecomes large.

2.2. Weighting coecient, w

One of the long-term goals of this work is to develop a method to incorporate DRTs into LSM's. Several studies have determined ``optimal'' view angles, which satisfy various criteria for sensible heat ¯ux calculation [2,4,8,13,28]. One rarely has control over the view angle of the sensor. In fact, from scanning sensors a wide range of zenith view angles are presented in any given surface temperature image. Therefore, the ability to accommodate DRTs measured from arbitrary view an-gles is important. The relationship between a DRT and

a given canopy temperature pro®le depends on the sensor view angle, and on the density and structure of the vegetation. These factors may be combined through

the temperature-weighting coecient, w, which appears

in (4) and is discussed below.

Brutsaert and Sugita [2] employ the approximation that DRT can be represented as a linear combination of the foliage skin temperature at the top and bottom of the canopy

hsˆwhfh‡ …1ÿw†hfg; …7†

where hs is the DRT measured at a known view angle.

The weighting coecient,w, depends on canopy density,

architecture, and skin temperature pro®le shape (i.e. the

decay coecient, b), as well as on the view angle of the

sensor.

Similarly in dual-source models [5,11], the DRT is taken to be a combination of the soil surface

tempera-ture,hsoil, and of the canopy temperature,hveg, assumed

to be isothermal, weighted by their respective fractions

fsoil and 1ÿfsoil visible to the sensor,

hsˆfsoilhsoil‡ …1ÿfsoil†hveg: …8†

More correctly, the radiances corresponding to these

temperatures should be added, but the linearization is a

good approximation provided hsoil andhveg are not too

di€erent.

For anisothermal canopies, (8) may be modi®ed by integrating through the canopy,

hsˆfsoilhsoil‡

where f is the fraction of the viewer ®eld not yet

ob-scured by foliage at the depthfinto the canopy;hf(f) the

foliage skin temperature given by (6).

For a vertically uniform canopy, the fractionfmay be

given by Beer's law,

f…f† ˆ exp

ÿg0LAIf=cosm ; …10†

whereg0is the leaf angle distribution parameter; LAI the

leaf area index;mthe sensor zenith view angle. Friedl [5]

estimatedg0usingg0ˆcos (ul), whereulis the mean leaf

inclination angle measured from a horizontal plane. For

grasslands and grass-like crops such as wheat,ulwill be

taken as 65°. Values of ulfor these and other types of

vegetation may be obtained from Gates [7]. The fraction

fsoil, which appears in (9) may be obtained by settingf

equal to 1 in (10).

For canopies whose foliage density is non-uniform in

the vertical direction, LAI*f may be replaced by

Rf

0a…f†df, where a(f) is a function which describes the

vertical distribution of canopy density as in Matsushima and Kondo [13].

Crago [4] obtained an expression for the weighting

(5)

(10), and setting the result equal to the right-hand side of

This requires the assumption that hsoilˆhfg (i.e., the

soil surface temperature is equal to the foliage skin temperature deep within the canopy at the limit of the

exponential canopy model (6), wheref! 1). This is a

reasonable assumption since (6) can even accommodate a canopy which is isothermal with a discontinuity be-tween the soil temperature and the foliage temperature,

by settingbˆ0, and lettinghfhbe the canopy

tempera-ture andhfg be the soil temperature.

Eqs. (4), (5) and (11) provide the information

re-quired for the sensitivity analysis of z0h. It is worth

noting that for the purpose of modelingz0h with (4), the

actual values of hfg and hfh are unimportant, but only

their di€erence.

2.3. Pro®le ®tting procedure

In (4),hfhis the temperature at the top of the canopy,

andhfg refers to the asymptotic limit of the exponential

pro®le as f (the normalized depth into the canopy

measured from the top downward) becomes large, i.e.,

below the bottom of the canopy.hfh,hfg and b are

un-knowns for each pro®le. We tookhfhdirectly from DRT

measurements at the ``top'' of the canopy, which we

describe more fully later. We estimated hfg and b by

®nding their values that minimized the sum of squared

di€erences between measured, hm(fi), and modeled,

h(fi), temperature pro®les

We accomplished this by the standard linear least squares procedure in which we took the partial

deriva-tive of (12) with respect tohfgafter substituting the

right-hand side of (6) forh(fi), set the result equal to zero, and

solved forhfg

hfgˆ

measurement heights in each pro®le.

Following a similar procedure for b produces an

implicit equation for b. Therefore, we linearized the

problem by rearranging (6) and taking the natural log-arithm.

If we set the partial derivative with respect tobof the

sum of the squared di€erences of the right- and

left-hand sides of (14), equal to zero, we can solve explicitly

for b

Eq. (15) may be substituted into (13) to eliminate b.

The resulting equation is an implicit equation for hfg

since bin (15) is a function ofhfg throughYi

hfgˆg…hfg†: …16†

Next, we de®ned a functionh…hfg† ˆhfgÿg…hfg† ˆ0

and found the values of hfg, which caused h(hfg) to be

zero. The exponentbis then calculated with (15). These

values ofhfgandbare those which minimize the sum of

squared di€erences between measured and modeled temperatures for each pro®le.

In addition, we imposed the constraints: (1) b>0;

and (2) hfg…ti†>hfg…tj† when hfh…ti†>hfh…tj†, where ti

andtjare the times of two di€erent pro®les. The ®rst of

these constraints is required in order for h(f) to

ap-proach hfg as f becomes large. The second constraint

was imposed on the basis of observations, which showed that pro®les tended not to overlap and that successive pro®les increase in curvature and steepness as tem-perature increases when averaged over a suitable time period on clear days.

In the following sections, we describe the data set used, and present results of ®eld observations of direc-tional radiometric canopy temperature pro®les from

which we obtain values for hfg,hfh, and b, and a

sensi-tivity analysis.

3. Data set

The data used here came from an experiment carried out from June 19 to July 9, 1997 in a hay ®eld north of

Boulder, Colorado, located at latitude 40° 60 52.800 N

and longitude 105°15051.600W. During this experiment,

DRT measurements of canopy temperature pro®les were collected. The data used here were from the com-pletely cloudless morning of July 1. The grass was characterized by a dense lower story 46 cm tall with

sparse individual stocks extending up to 87 cm. The 15°

(6)

starting at 7:00 and 9:10 a.m., respectively. From these measurements, more than 18 500 values were consoli-dated into 30-min averages to which we refer in the text using the ending time of the 30-min (e.g., we refer to the average from 7 to 7:30 a.m. as ``pro®le 7.5''). A few of the time periods were shorter than 30 min (6:18±6:30; 7:04±7:30; 9:00±9:10; and 11:30±11:50). We still refer to these by the ending time of the half-hour period. Table 1 contains the values of these 30-min pro®les as a function of time and height above the ground.

4. Results and discussion

4.1. General observations

The 30-min averaged DRT pro®les appear in Fig. 2. Several important features show up here. First, the pro®les begin nearly isothermal at 6:30 a.m., and the mean pro®le temperature increases and the temperature gradient steepens as the morning progresses. Second, the

sparse upper canopy remains nearly isothermal

throughout the morning, and most of the gradient is concentrated in the dense lower canopy.

Since z0h given by (4) depends on the di€erence in

temperatures between the bottom and top of the

cano-py, …hfgÿhfh†, we have plotted …hf…7 cm† ÿhfh† from

each 1-min pro®le as a function of time in Fig. 3, to get

an idea of how …hfgÿhfh†might behave. hfh is the

av-erage of all DRTs measured at and abovezˆ42 cm, and

hf(7 cm) is the temperature measured closest to the

ground, atzˆ7 cm. Hereafter,hf(7 cm) will be denoted

hf7. The vertically averaged temperature,hfh, was used to

represent the top of canopy temperature because the upper canopy was relatively sparse, approximately iso-thermal, and if the pro®les exhibit an exponential shape, this occurs in the lower part of the canopy at and below 42 cm, as is clear in Fig. 2.

Table 1

DRT values of 30-min averaged pro®les in°C as functions of time and height above the ground surface

Height (cm)

Time (h)

6.5 7 7.5 8 8.5 9 9.5 10 10.5 11 11.5 12

84 12.7 14.4 17.9 18.7 19.3 20.0 20.2 21.7 22.1 22.7 23.0 23.6

77 12.3 14.2 17.5 18.3 19.0 19.8 20.1 21.6 21.9 22.6 22.8 23.5

70 11.9 13.9 17.2 18.0 18.8 19.7 20.1 21.7 21.9 22.6 22.7 23.5

63 11.6 13.7 17.1 17.9 18.7 19.7 20.2 21.9 22.0 22.7 22.9 23.8

56 11.2 13.6 16.9 17.9 18.7 19.7 20.3 22.0 22.3 23.0 23.0 23.9

49 10.8 13.5 16.8 17.8 18.6 19.6 20.2 22.2 22.5 23.4 23.6 24.2

42 10.6 13.1 16.6 17.5 18.5 19.5 20.2 22.3 22.7 23.8 24.2 25.1

35 10.6 12.8 16.7 18.1 19.1 20.3 20.7 23.1 24.0 25.0 25.9 27.2

28 10.5 12.6 17.1 18.7 19.7 20.9 21.5 24.1 25.6 27.1 28.0 29.4

21 10.5 12.5 16.8 18.9 20.2 21.8 22.4 25.4 26.9 28.5 29.7 31.3

14 10.9 12.8 16.6 19.1 20.8 22.3 23.2 26.6 28.0 29.6 31.1 33.1

7 11.4 13.2 16.4 19.1 21.4 22.8 23.9 27.4 28.8 30.5 32.2 34.0

Fig. 2. A 30-min averaged DRT pro®les. Numbers indicate range of times (HHMM) included in each pro®le.

Fig. 3. Evolution of bottom to top canopy temperature di€erences, hf7ÿhfh, for each one-minute pro®le, and for 30-min averages with

(7)

The solid line in Fig. 3 connects the values of

…hf7ÿhfh† taken from the 30-min averaged data. In

addition, since we are interested in the diurnal

pro-gression of…hf7ÿhfh†and its causes, we have plotted the

solar zenith angle (dashed line) versus time in Fig. 3. This illustrates qualitatively, that as the sun rises higher in the sky (decreasing zenith angle), it penetrates more deeply into the canopy and causes signi®cant warming of the lower canopy. The upper canopy remains cooler due to exposure to the wind.

The mean di€erence…hf7ÿhfh†is zero at around 7:30

a.m. It then increases relatively steadily throughout the

morning to a mean di€erence of about 10°C (‹2°C) by

11:30 a.m. Solar noon occurs at this location at

ap-proximately 12:58 MDST. If …hf7ÿhfh† continued to

increase at the same rate shown in Fig. 3, one might

expect…hf7ÿhfh†to increase by another 3°C to a value

of 13°C by solar noon at which time the solar zenith

angle would be approximately 17°. This increase in

…hf7ÿhfh† would correspond to a decrease in z0h

throughout the morning according to (4), other factors remaining the same.

4.2. Fitted pro®les

The derivation of (4) in Brutsaert and Sugita [2] as-sumes an exponential temperature pro®le. To evaluate the realism of this assumption, we ®tted exponential curves described by (6) to the 30-min average DRT pro®les, using the least squares procedure described earlier. We used the average of the temperatures at and

above 42 cm forhfh for the reasons explained above.

hfg and b converged to solutions which satis®ed

h(hfg)ˆ0 and the imposed constraints for ®ve of the

pro®les (#s 8, 10.5, 11, 11.5, and 12). The other cases failed for the following reasons:

1. Pro®les 7.5, 8.5, and 9.5 violated constraint 2. Due to

the small range inh(f), curvature in the pro®le is

ob-scured by the scatter in the data. The optimization

procedure tends toward a straight line for h(f) over

the applicable range of f. This requires b!0 and

hfg! 1, which caused a divide by zero error in (13).

2. Pro®les 9 and 10 violated constraint 2. As in (1) above, the curvature is slight. The optimization

pro-cedure found a solution buthfg was larger than that

for later pro®les.

3. Pro®les 6.5 and 7 reveal internal minima within their pro®les so that the least squares solution becomes a vertical line through the average value of the lower

®ve temperatures. This forcesbto become excessively

large in order forh(0) to equalhfh.

In order to ®t an exponential pro®le to those pro®les which failed to converge and/or satisfy the constraints,

we examined the temporal trend in…hfgÿhfh†for the ®ve

pro®les for which the solution converged and satis®ed the constraints, and compared the trend to that for

…hf7ÿhfh†. We illustrate the comparison in Fig. 4.

Considering the solid squares and their corresponding

diamonds, it appears that …hfgÿhfh†is roughly

propor-tional to …hf7ÿhfh†. This is physically realistic since (1)

both …hfgÿhfh†and …hf7ÿhfh† should become 0

simul-taneously when the pro®le is isothermal, and (2) each successive pro®le appears to envelop the previous pro-®les and increase in steepness and curvature in Fig. 2. The average value of the ratio of …hfgÿhfh†=…hf7ÿhfh†

was 1.46. We determined …hfgÿhfh† for the other

pro®les by multiplying their corresponding …hf7ÿhfh†

value by this ratio. The results are shown by the hollow square symbols in Fig. 4.

For each pro®le, we used these values of…hfgÿhfh†to

determine hfg, and then determinedbby means of (15).

Doing so gives the value ofbthat minimizes the sum of

squared errors associated with the linearized

tempera-ture pro®le equation (14) for the speci®ed value of hfg.

For each time period, Table 2 presents the values ofhfg,

b …hfgÿhfh†, and the root-mean-square error (RMSE)

between measured and modeled temperature pro®les. The measured (symbols) and modeled (lines) pro®les determined from (6) using the data in Table 2 are shown

in Fig. 5. The resulting values of b are all very similar

and range between about 1.1 and 1.4 with an average value of 1.26, as shown in Fig. 6. The relatively constant

nature ofbis one of the important ®nding of this paper.

It is important to note that the RMSE did not

in-crease substantially for those pro®les for whichhfgandb

were estimated based on the ration method. In fact, the

RMSE was relatively insensitive tohfg. We illustrate this

in Fig. 7. The solid line in panel (a) shows the RMSE as

a function ofhfg for pro®le 10. In decreasing the global

optimum value of hfgˆ66°C by 26±40°C, RMSE only

increases from about 0.11±14°C. Thereafter, the slope of

RMSE increases, but in going tohfgˆ30.0°C (the value

Fig. 4. Comparison between the temporal trends inhf7ÿhfh(diamond

symbols) and hfgÿhfh (square symbols). The solid square symbols

denotehfgÿhfhvalues determined by the least squares pro®le ®tting

procedure. The hollow square symbols denotehfgÿhfhvalues

(8)

obtained by the ratio method), RMSE only increases to

0.34°C. Although this is larger than the minimum value

by a factor of 3, it still represents only 3.75% of the

di€erence …hfgÿhfh†, which is certainly an acceptable

error. The signi®cance of this is that determination ofhfg

by the ratio method did not substantially increase the error in modeling the measured pro®le.

In contrast, the pro®les 8, 10.5, 11, 11.5 and 12, which satis®ed the constraints in our original optimi-zation run, exhibited much more signi®cant changes in

RMSE for small changes inhfg. The solid line in panelb

of Fig. 7 shows that an increase in RMSE for pro®le 12

from the global minimum to a value of 0.3°C only

allows a range inhfgfrom about 35.5°C to 44°C. This is

much smaller than the range 30°C to +1 allowed for

pro®le 10 for the same change in RMSE, and indicates

that the certainty in the value ofhfgfor pro®les 8, 10.5,

11, 11.5, and 12, which satis®ed the constraints in the original optimization run is much higher than for the other pro®les.

4.3. Sensitivity analysis

To demonstrate the behavior ofz0h, we plotted, on a

logarithmic scale in Figs. 8(a±d), the ratio z0h=z0hi

de-termined from (4) and (5) as a function of several vari-ables. Values of the variables for the base scenario and their ranges are given in columns 2 and 3 of Table 3.

Fig. 7. Value of root mean square error (RMSE) as a function ofhfg.

The dashed line represents the locally optimum value of b corre-sponding tohfg. The RMSE for any pair ofhfg,bvalues is given by the

value of the RMSE line directly above the pair. Panel (a) corresponds to pro®le 10; panel (b) corresponds to pro®le 12.

Fig. 5. Measured (symbols) and modeled (lines) temperature pro®les for lower half of the canopy, where the exponential pro®le appears to exist.

Fig. 6. Illustration of approximate constancy of the coecient,b, in the exponential pro®le model (6).

Table 2

Values of variables used to model the canopy temperature pro®le by means of (6), and the RMSE between measured and modeled values

Time (h) hfg(°C) b hfgÿhfh

(°C)

RMSE (°C)

6.5 9.98 1.07 )1.61 0.530

7 11.9 1.18 )1.87 0.455

7.5 16.1 1.18 )1.04 0.163

8 19.7 1.44 1.63 0.137

8.5 22.5 1.19 3.74 0.255

9 24.2 1.29 4.48 0.208

9.5 25.6 1.24 5.39 0.322

10 30.0 1.29 8.03 0.339

10.5 32.9 1.16 10.7 0.050

11 34.3 1.31 11.4 0.099

11.5 36.3 1.39 13.2 0.038

(9)

Panel (a) of Fig. 8 shows that z0h=z0hi decreases

lin-early on a logarithmic scale as a function of increasing

temperature di€erence …hfgÿhfh†. This is also clear

based on simple inspection of (4). Panel (a) also shows the e€ects of increasing LAI, which are not as readily apparent from (4). An eightfold increase in LAI from

0.5 to 4 causes an increase inz0h=z0hifrom about 0.024 to

0.31 for the base scenario. The incremental increase in

z0h=z0hi decreases for each additional unit increase in

LAI (e.g., there is a bigger change in going from LAIˆ1

to 2, than from 2 to 3, etc.).

Althoughz0h is non-linearly related tou, due to the

presence of u in (4) in the numerator and in

C2 ‰ˆf…Ctf…Re…u†††Š, the direct in¯uence of u in the

numerator dominates, so that z0h=z0hi is essentially

linearly related to u on a logarithmic scale. z0h=z0hi

decreases from about 0.66 to about 0.02 for a change

in u from 0.1 to 1 m/s for the base scenario (not

shown).

The ratioz0h=z0hi is not very sensitive tob. Panel (b)

shows that a change in b from 0.5 to 1.5 produces a

negligible change even for …hfgÿhfh† ˆ15°C. This is

signi®cant given the narrow range ofbobserved for this

data set. Furthermore, even a large change inbfrom 1.5

to 7 only increases z0h=z0hi from around 0.26 to 0.4 for

the base scenario. Crago [4] demonstrates that the

weighting coecient, w, is very sensitive tob, however,

apparently this sensitivity does not propagate its way

through to z0h. Panel (c) shows that z0h=z0hi becomes

more sensitive tobas LAI decreases.

Finally, the e€ect of zenith view angle for a sensor located above the canopy is shown in panel (d). View

angle a€ects z0h=z0hi by means of changes in the

weighting coecient, w given by (11), which relates

DRT to hfg andhfhthrough (6). Note here that, unlike

LAI, a unit increase in view angle creates successively

larger increases in z0h=z0hi as view angle increases. For

the base scenario, a view angle of 71.39°causesz0h=z0hi

to equal 1 for all values of …hfgÿhfh†. Others have

commented on the fact that various models and

obser-vations suggest that a large view angle causes z0h to

Fig. 8. Results of sensitivity analysis for the base scenario and variable parameter values indicated in Table 3.

Table 3

Values of variables used for base scenario of sensitivity analysis and ranges of values allowed to vary in Figs. 8(a±d)

Variable Base scenario Range

hfgÿhfh 5°C 0±15°C

B 1 0.5±7

LAI 3 0.5±4

u 0.4 m sÿ1

m 20° 0±70°

H 0.42 m

A 3

z0m h/8

d0 2/3h

Ctf 0.0715

H 150 W mÿ2

(10)

behave as though the canopy were isothermal [2,4,8,13,28].

5. Summary and conclusions: signi®cance to modelingz0h

To summarize these results, it is useful to review some

of the background on the scalar roughness,z0h. First, a

parameterization for z0h appears in all meteorological

weather models [27], therefore improvements to it will have broad impacts. Secondly, many parameterizations

forz0hhave been proposed. Research indicates thatz0his

neither constant nor is it proportional to the momentum

roughness, z0m. Furthermore, the roughness Reynolds

number does not explain all of its diurnal variability. Some of the diurnal variability may be due to vertical skin temperature gradients that form within vegetation as the sun rises higher in the sky and penetrates more deeply into the canopy. Brutsaert and Sugita [2] derived an expression to account for the in¯uence of

tempera-ture gradients onz0h. Their derivation assumed an

ex-ponential pro®le shape for the temperature gradients given by (6). This assumption limits the applicability of their method to canopies with a relatively uniform ver-tical structure. In this paper, we present data to evaluate this assumption and examine the semi-diurnal behavior

of several variables,b,hfg, andhfhwhich are important

parameters in the exponential temperature pro®le (6). The speci®c conclusions from this work are listed below:

1. The exponential decay coecient,b, may be

approx-imately constant for a given canopy at a given growth stage, although it is likely to be a function of LAI

(Fig. 6). This, together with the fact that z0h does

not appear to be extremely sensitive tob (Fig. 8(b)),

suggests that for a given canopy,b may not have a

signi®cant in¯uence on z0h. Furthermore, b may be

able to be estimated for a given canopy on the basis of a few pro®le measurements when solar zenith angle

is small so that the curvature and range ofh(f) in the

pro®le is large.

2. …hfgÿhfh†appears to increase in a consistent fashion

on clear days as solar zenith angle decreases (Figs. 3

and 4). This behavior and the ®nding thatb is

rela-tively constant may provide useful constraints for

es-timating…hfgÿhfh† and b from time series of DRTs

measured from multiple view angles.

3. If we extrapolate the rate of increase in…hfgÿhfh†to

solar noon (1 p.m. MDST), it would attain a value

of20°C. Thus for this canopy/location/time of year,

a rough estimate for the diurnal range of…hfgÿhfh†is

0±20°C.

Although these conclusions do not provide enough in-formation to apply Brutsaert and Sugita's [2]

par-ameterization for z0h operationally, they do provide

some simpli®cations that will be helpful to evaluate the

diurnal variability ofz0h.

Acknowledgements

This material is based upon work partially supported by the National Science Foundation under Grant No. EAR-9903117 and by the National Aeronautics and Space Administration under Grant No. NAG5-8534.

References

[1] Brutsaert WH. Evaporation into the atmosphere: theory, history and applications. Hingham, MA: Reidel; 1982.

[2] Brutsaert W, Sugita M. Sensible heat transfer parameterization for surfaces with anisothermal dense vegetation. J Atmos Sci 1996;53:209±16.

[3] Chen F, Janjic Z, Mitchell K. Impact of atmospheric surface layer parameterization in the new land-surface scheme of the NCEP mesoscale Eta numerical model. Boundary-Layer Meteorol 1997;85:391±421.

[4] Crago R. Radiometric and equivalent isothermal surface tem-peratures. Water Resour Res 1998;34:3017±23.

[5] Friedl JA. Modeling land surface ¯uxes using a sparse canopy model and radiometric surface temperature measurements. J Geophys Res 1995;100:25435±46.

[6] Garratt JR, Francey RJ. Bulk characteristics of heat transfer in the unstable baroclinic atmospheric boundary layer. Boundary-Layer Meteorol 1978;15:399±421.

[7] Gates DM. Biophysical ecology. New York: Springer; 1980. [8] Huband NDS, Monteith JL. Radiative surface temperature and

energy balance of a wheat canopy, 1, comparison of radiative and aerodynamic temperatures. Boundary-Layer Meteorol 1986;36:1± 17.

[9] Kimes DS, Idso SB, Pinter PJ Jr., Reginato RJ, Jackson RD. View angle e€ects in the radiometric measurement of plant canopy temperatures. Remote Sens Environ 1980;10:273±84.

[10] Kustas WP, Choudhury BJ, Moran MS, Reginato RJ, Jackson RD, Gay LW, Weaver HL. Determination of sensible heat ¯ux over sparse canopy using thermal infrared data. Agric For Meteorol 1989;44:197±216.

[11] Kustas WP, Norman JM. A two-source approach for estimating turbulent ¯uxes using multiple angle thermal infrared observa-tions. Water Resour Res 1997;33:1495±508.

[12] Lagouarde JP, Kerr YH, Brunet Y. An experimental study of angular e€ects on surface temperature for various plant canopies and bare soils. Agric For Meteorol 1995;77:167±90.

[13] Matsushima D, Kondo J. A proper method for estimating sensible heat ¯ux above a horizontal±homogeneous vegetation canopy using radiometric surface observations. J Appl Meteorol 1997;36:1696±711.

[14] Norman JM, Becker F. Terminology in thermal infrared remote sensing of natural surfaces. Agric For Meteorol 1995;77:153±66. [15] Pan H-L, Mahrt L. Interaction between soil hydrology and

boundary-layer development. Boundary-Layer Meteorol 1987;38:185±202.

[16] Prevot L, Brunet Y, Paw U KT, Seguin B. Canopy modeling for estimating sensible heat ¯ux from thermal infrared measurements. In: Proceedings of the Conference on Thermal Remote Sensing of the Energy and Water Balance over Vegetation in Conjunction with Other Sensors; 1993.

[17] Qualls RJ, Brutsaert W. E€ect of vegetation density on the parameterization of scalar roughness to estimate spatially distrib-uted sensible heat ¯uxes. Water Resour Res 1996;32:645±52. [18] Qualls RJ, Hopson T. Combined use of vegetation density,

(11)

[19] Sellers PJ, Mintz Y, Sud YC, Dolcher A. A simple biosphere model (SiB) for use within general circulation models. J Atmos Sci 1986;43:505±31.

[20] Smith JA, Goltz SM. Updated thermal model using simpli®ed short-wave radiosity calculations. Remote Sens Environ 1994;47:167±75.

[21] Smith JA, Chauhan NS, Schmugge TJ, Ballard JR., Jr. Remote sensing of land surface temperature: the directional viewing e€ect. IEEE Trans Geosci Remote Sensing 1997;35:972±74.

[22] Stewart JB, Kustas WP, Humes KS, Nichols WD, Moran MS, de Bruin HAR. Sensible heat ¯ux±radiometric temperature relation-ship for eight semiarid areas. J Appl Meteorol 1994;33:1110±7. [23] Sugita M, Brutsaert WH. Regional surface ¯uxes from remotely

sensed skin temperature and lower boundary layer measurements. Water Resour Res 1990;26:2937±44.

[24] Sugita M, Kubota A. Radiometrically determined skin tempera-ture and scalar roughness to estimate surface heat ¯ux, Part II: performance of parameterized scalar roughness for the

determi-nation of sensible heat. Boundary-Layer Meteorol 1994;69:397± 416.

[25] Sun J, Mahrt L. Determination of surface ¯uxes from surface radiative temperature. J Atmos Sci 1995;52:1096±106.

[26] Van de Griend AA, van Boxel JH. Water and surface energy balance model with a multilayer canopy representation for remote sensing purposes. Water Resour Res 1989;25:949±71.

[27] Verhoef A, de Bruin HAR, van den Hurk BJJM. Some practical notes on the parameter kBÿ1 for sparse vegetation. J Appl

Meteorol 1997;36:560±72.

[28] Vining RC, Blad BL. Estimation of sensible heat ¯ux from remotely sensed canopy temperatures. J Geophys Res 1992;97:18951±4.

Referensi

Dokumen terkait

Tingkat pelaksanaan PBL yang sedang ini, menunjukkan bahwa (a) mahasiswa telah membangun ilmu pengetahuannya, namun belum maksimal, (b) kontrol proses belajar telah berada pada

Dengan mengetahui pengaruh ekspresi p53 dan HIF1 terhadap peningkatan kadar laktat jaringan nasofaring pada pasien KNF WHO tipe III, diharapkan dapat diketahui pola produksi

bahwa berdasarkan pertimbangan sebagaimana dimaksud dalam huruf a dan untuk melaksanakan ketentuan Pasal 4 ayat (4) Undang-Undang Nomor 19 Tahun 2003 tentang Badan Usaha Milik

Penerapan inkuiri moral berbasis nilai-nilai kearifan lokal minangkabau “alam takambang jadi guru” untuk pembentukan karakter siswa.. Universitas Pendidikan Indonesia |

Saran bagi Guru BK, hendaknya dapat mengoptimalkan layanan informasi sebagai salah satu alternatif untuk membantu siswa dalam meningkatkan motivasi belajar.Bagi sekolah,

dalam suatu organisasi, termasuk pemerintah daerah yang harus mengelola APBD. dimana volume transaksinya dari tahun ke tahun menunjukkan peningkatan

web ini adalah untuk memudahkan pihak sekolah dalam memberi informasi kepada masyarakat. dan user/pengguna dapat meng-update isi dari informasi yang terdapat pada web

[r]