Sie sind auf Seite 1von 10

ELSEVIER

AGRICULTURAL
AND
FOREST
METEOROLOGY
Agricultural and Forest Meteorology 88 (1997) 47-56

Model for evaporation, moisture and temperature of bare soil:


calibration and sensitivity analysis
Gunnel Alveniis

*, Per-Erik

Jansson

Department qf Soil Sciences, Swedish University of Agricultural Sciences, P.O. Box 7014, Vppsala S-75007, Sweden

Received 30 December 1996; received in revised form 20 June 1997; accepted 23 June 1997

Abstract
A modelling approach for predicting soil surface temperature and soil evaporation is presented. The procedure is based
on the equations for heat flow at the soil surface and includes vapour diffusion and a semi-empirical correction function for
the surface vapour pressure. The effects of changes in three important model parameters were studied by means of multiple
model simulations. The first parameter determines the steepness of the water potential gradient close to the surface. The
second parameter is the water vapour enhancement factor and the third one limits the lowest possible hydraulic conductivity
during drying. Measurements of soil water content and soil temperature in a bare sandy loam were used to evaluate the
models behaviour. Vapour pressure at the soil surface was found to be substantially lower than saturated vapour pressure at
topsoil moisture potentials as high as - 100 hFa. The difficulty to distinguish between vapour and liquid water flow at low
moisture contents was demonstrated. Results from the temperature tests indicated enhanced vapour diffusion and a probable
value of the diffusion tortuosity coefficient close to 1.0, whereas a value close to 0.7 was more likely according to the soil
water contents and calculated evaporation. 0 1997 Elsevier Science B.V.
Keywords: Energy balance; Soil surface; Vapour pressure; Soil temperature; Soil water content; Model; Simulation; Soil evaporation;
Vapour flow

1. Introduction
Evaporation from a bare soil surface is a complex
process including multi-phase transport of soil water
to the evaporating surface. This results in a vertical
redistribution
caused by simultaneous
liquid and
vapour water flows. In addition, evaporation depends
not only on soil properties but also on atmospheric

Corresponding author. Fax: + 46-18-672795; e-mail:


gunnel.alvenas@mv.slu.se
Fax: +46-18-672795; e-mail: per-erik.jansson@mv.slu.se
??

conditions. Although, during the few last decades,


much effort has been spent on describing the processes involved, there are still problems left to be
solved such as those concerned with the humidity
conditions of a non-saturated soil surface. A number
of authors (e.g., Deardorff, 1978; Barton, 1979; Yasuda and Toya, 1981; Alveniis et al., 1986; Kondo et
al., 1990) have presented soil evaporation models
with different
parameterization
schemes for the
evaporating surface, and there are many divergent
opinions as to which methods are the most correct
and useful (e.g., Kondo et al., 1990; Lee and Pielke,
1992; Mihalovik et al., 1993; DekirZ et al., 1995).

0168-1923/97/$17.00 0 1997 Elsevier Science B.V. All rights reserved.


PZZSO168-1923(97)00052-X

48

G. Alueniis, P-E. Jansson /Agricultural and Forest Meteorology 88 (1997) 47-56

For a saturated

surface, the soil evaporation,

E,

can be written as

where p is the density of air, cp is the specific heat


of air, h is the latent heat of vaporization, y is the
psychrometer constant, e * (T,) is the saturation vapor
pressure at the surface temperature, T,, e is the air
vapor pressure at reference level and ra is the aerodynamic resistance to water vapor exchange. When
an earlier saturated soil dries, the free water in larger
pores evaporates first. The remaining water is retained mainly in smaller soil pores by strong capillary forces, and the vapor pressure of the air in
equilibrium
with the pore water will therefore be
lower than in the air close to a free water surface.
Two common ways to compensate for the influence
of soil water content on surface vapor pressure are
the CYand p methods:

(2)
(3)
where (Y and p, often called the soil moisture
availability factors, are functions of soil water content that compensate for the changes in soil surface
humidity during drying. Comprehensive overview of
(Y and p formulations occurring in the literature can
be found in Mafouf and Noilhan (1991), Lee and
Pielke (1992), Mihalovid et al. (1993), and Deki6 et
al. (1995).
From thermodynamic
laws, Philip (1957) derived
an expression for the relative humidity of air in
equilibrium with the water in the soil pore, h, as:
h=exp($M,g/R(T+273.15))

when the upper layer is dry. A prerequisite for the


equation to be valid is that the air close to the
pore-water surface is in equilibrium with the pore
water. However, in a drying soil, where vapor is
continuously transported to the atmosphere, the vapor pressure of air adjacent to the evaporating water
surface will not be in equilibrium
with the liquid
water in the pore, i.e., h does not represent the
specific humidity at the soil surface, cy. Moreover,
using Philips formula and (Y= h, the air close to the
water in the pore will be saturated ((Y = 1) at very
low soil water contents (Kondo et al., 1990) and will
not start to decrease until the water content has
dropped far below the permanent wilting point (Lee
and Pielke, 1992). This will lead to an overestimation of soil evaporation (MihaloviC et al., 1993).
Kondo et al. (1990, 1992) tried to overcome the
deficiency in Philips equation by introducing
an
empirical resistance to the transport of water vapor
from the soil pores to the soil-atmosphere
interface.
Lee and Pielke (1992) proposed a new formulation
where the soil moisture availability factor, /3, starts
to decrease once the soil water content drops below
field capacity. The introduction of field capacity as a
common reference point eliminated the dependence
on soil texture. These formulas predict a smoother
transition in the soil surface specific humidity between wet and dry soil states (Mihalovic
et al.,
1993). A similar formulation for the p method was
earlier presented by Deardorff (1978).
In this paper, we present a method to correct
Philips equation, where the surface humidity is calculated based on the soil water potential in the
topsoil layer and a dynamic estimate of the water
balance at the surface. Thereby, we account for
non-equilibrium
effects caused by rapid moisture
fluctuations close to the surface.

(4)

where $ is the soil water potential at the surface,


M, is the molecular
weight of water, g is the
acceleration due to gravity and R is the gas constant
for water vapor at temperature T. Philips expression
has been used to simulate surface air humidity, h,,
by several authors (e.g., McCumber and Pielke, 1981;
Camillo et al., 1983; Nappo, 1975), but as Kondo et
al. (1990) and Lee and Pielke (1992) suggest, Hq. (4)
may be invalid close to a soil surface, especially

2. Materials

and methods

2.1. Model theory


Heat flow in the soil, G, can be calculated
sum of soil heat conduction, the convection
due to water flow and latent heat convection
by water vapor flow. With the indices w

as the
of heat
caused
and v

G. Alweds, P-E. Jansson/Agricultural and Forest Meteorology 88 (1997) 47-56

49

representing water and vapor, the heat flow equation


can be written as:

(11)

where q is flow, k is conductivity, T is the temperature, z is depth, c, is the specific heat of water and
A is the heat of vaporization.
Soil water flow is comprised of saturated flow
according to Darcys law as generalized for unsaturated flow by Richards (1931) and water vapor flow,
q,, as follows:
4, = k

d+
-z+1

+4v

(6)

where I/J is the soil matric potential. Water vapor


flow in the model is mainly based on Ficks law of
diffusion, modified for soils:

and pa is the density of air, cp the specific heat of


air, y the psychrometer constant and k, the heat
conductivity of the soil. ra is a resistance to transport
of heat and water vapor between the surface and the
atmosphere, which depends on wind speed, surface
roughness and stability of the atmosphere. The vapor
flow, q,, is calculated based on differences in soil
vapor concentrations, C, between the surface and the
centre of a top-soil layer, AZ, according to:
4, =

dv,D,fa

cs- csoil

(12)

AZ
2

The surface vapor pressure, e,, is expressed as:


*M,ge,
e, = e*(T,)e(R(r+273.15))

where D, is the diffusion coefficient for free air, f,


is the soil air content, C, is the water vapor concentration and d, is a lumped parameter accounting for
tortuosity and the enhancement of vapor transfer
observed in measurements as compared with theory
(i.e., Philip and de Vries, 1957; Cary, 1963).
Soil evaporation, E, can be calculated by solving
of the heat flow equations at the surface, i.e.,

(13)

where e *(T,) is the saturated vapor pressure at the


surface temperature, $ is the mean soil moisture
potential in the top-soil layer, g is the gravity constant, and R the gas constant at temperature T,. e, is
an empirical correction factor which compensates for
the difference between the mean soil moisture potential in the top-soil layer and the soil moisture potential at the surface, defined as:
e, = 10(-ss+g)

R,=hE+H+G

where R, is the net radiation at the surface, A is the


heat of vaporization, H is the sensible heat flux from
the surface to the atmosphere and G is the soil heat
flux. The exchange processes are governed by different temperature, T, and vapor pressure, e, conditions
in the soil and atmosphere. With the indices a, soil,
and s representing air, soil and soil surface respectively, the processes can be written as:

H= pacp

(14)

(8)

(K-T,)
ra

where 8, is the surface water balance based on the


difference between precipitation, P, and evaporation
and soil vapor flux. 8, is only allowed to vary
between -2 and 1 mm according to:
8,(t) =max( -2,min(1,6,(t+(P-E-q,)At)

1)
(15)

In combination with S,, I& in Eq. (14) is a


parameter the value of which determines the steepness of the soil water potential gradient from the
middle of the top layer to the surface. +g = 0 implies
that there is no difference between the water potential at the surface and that in the top-soil layer.
I,$ = 1 implies that the water availability at the sur-

50

G. Alveniis, P-E. Jansson /Agricultural and Forest Meteorology 88 (1997) 47-56

face could decrease by up to two powers of magnitude during drying and increase by up to one power
of magnitude during wetting.
The precise predictions of soil surface evaporation
will thus include empirical parameter values for
vapour diffusion in the soil, d,, and the non-equilibrium coefficient, $s. In reality, the unsaturated
conductivity, k,, is also difficult to estimate under
dry conditions. To investigate the importance of the
unsaturated conductivity values at low water contents the k, equation by Mualem (1976) is limited
with a minimum value of unsaturated conductivity,
k,. Thus,
k, = max[ks( f)2@l*.

k-1

(16)

where k, is the saturated conductivity, +!I~is the air


entry pressure, h is the pore size distribution index
and n is a parameter accounting for pore correlation
and flow path tortuosity.
2.2. Field measurements
A number of different measurements were carried
out close to Uppsala (60 N) in central Sweden in a
small (2 X 2 m), bare plot with a 0.3-m homogeneous sandy loam layer overlying a coarse sand.
Besides the standard meteorological measurements at
the site (air temperature and humidity, radiation,
wind speed and precipitation) soil and soil surface
temperature and moisture were registered continu-

Soil water content (K of volume)

ously. Thin thermocouples (type T, copper-constantan) for soil temperature measurements were installed at l-cm depth intervals from 1 to 6 cm depth
in two profiles in the plot, while temperatures at
deeper levels (10, 20 and 30 cm) were measured
only in one profile. Soil water contents were measured with Time-Domain Reflectometry (TDR)
technique in two similar profiles at 2, 4, 6, 10, 20
and 30 cm depth, respectively. Miniature tensiometers were used for soil water tension measurements
at the same levels as the TDR probes in one profile.
Soil surface temperature was determined both with
four thin thermocouples and with an infrared thermometer scanning over the plot. Soil surface moisture was measured with a gas analyser on air sampled close to the soil surface. The plot was equipped
with a miniature net radiometer.
2.3. Soil properties

Soil texture from 0 to 30 cm depth was determined at 5-cm depth intervals for lumped soil samples at three different spots. The soil profile from 0
to 30 cm depth was homogeneous, consisting of
approximately 10% of clay, 13% of silt and 73% of
sand and with an organic matter content of 3-4%.
The bulk density increased from 1.40 g cme3 at O-5
cm depth to 1.63 g cme3 at 25-30 cm depth.
Small soil cores (5 cm high, 7 cm of diameter)
were sampled from the surface down to 30-cm depth,
with three replicates at each level, to obtain informa-

Soil water potential (hPa)

Fig. 1. (a) Soil water retention curves for depth intervals of O-5 and 5-30 cm. The curves are fitted to measured values (crosses and
squares) using the Brooks and Corey expression. (b) Unsaturated conductivity for O-30 cm depth calculated according to Mualem (lines)
and calculated from measurements (crosses = 3 cm, filled circles = 10 cm and squares = 20 cm depth).

G. Aluenti,

P-E. Jansson/Agricultural

tion on soil water retention properties (Fig. la).


Separate samples for determination of saturated conductivity were collected down to lo-cm depth. The
expression presented by Mualem (1976) was used to
calculate the values of unsaturated conductivity to be
applied in the simulations. Unsaturated conductivity
was also calculated using measured values of soil
water contents and tensions at different levels in the
profile (Fig. lb).
2.4. Sensitivity analysis

and Forest Meteorology

40

51

SOlI 5 em depth

(C)

30 20-

ii.
i,
9

10 -

0,

30 _

Volumetric water content

I-

Scmdepth

(%)

1:
12

The SOIL model (Jansson, 1991) was parameterized for 14 layers, the thickness of which increased
slowly with depth, starting with 2 cm layers in the
uppermost 8 cm of the profile. The meteorological
station at the site provided data for the driving
variables. Measured temperature and water content
were used as initial values. To represent a plain bare
soil surface the surface roughness was estimated to
be 0.001 m.
Three model parameters affecting soil surface
moisture and temperature were selected for a sensitivity test, namely the I,!+, k, and d,, parameters.
The first parameter governs the surface moisture
under both dry and wet conditions, while the importance of the two latter increases when the soil becomes dry. The I& parameter was increased in steps
of 0.4 from 0.0 to 2.0 which covers a wide range of
moisture gradients in the uppermost layer, allowing
the surface water availability to decrease up to lo4
times during drying. A commonly used value of the
d,, parameter in the diffusion equation is 0.66, but
higher possible levels, even above 1.0, have been
suggested (Philip and de Vries, 1957; Cary, 1963;
Cass et al., 1984). This enhanced vapour flow may
be caused by e.g., local temperature gradients or
evaporation/condensation
on different sides of thin
water films in the soil (Philip and de Vries, 1957).
An extensive overview over different explanations of
the enhancement phenomenon and a presentation of
vapor diffusion enhancement
factors occurring in
literature can be found in ten Berge (1986). In the
present test, the d,, parameter was increased from
0.1 to 1.0 in steps of 0.3 to cover both low-diffusion
scenarios and enhanced vapour diffusion. The minimun hydraulic conductivity,
k,, was increased tenfold in every step :from 10- to 10 mm/day,

Temperature

88 (1997) 47-56

27

30,

June
Fig. 2. Measured air and soil temperature
5-cm depth.

J&
and soil water content at

which represents both very low and high conductivities.


Data from a lCday-long
period in June-July
1994 were used in this study. Dry weather conditions
dominated during the entire period except for the
night between 29 June and 30 June, when 4 mm of
rain fell. In the early afternoon of July 7, the plot
was irrigated with 10 mm of water.
Simulated values of soil water content at depths
of 3, 5, 10,20, and 30 cm and temperature at depths
of 1, 5, 10, 20, and 30 cm were compared with
corresponding
measured values (Fig. 2), and root
mean square errors (RMSE) and coefficients of determination (r*) were calculated. Water contents at
the 3- and 5-cm depths were interpolated from measurements at 2- and 4-cm depths and at 4- and 6-cm
depths respectively.
The calculated RMSE and r*
values at different depths were used to evaluate the
effect of different parameter values on simulated
temperatures and water contents. Parameter combinations that gave rise to high RMSE values and low r*
values were discarded.

3. Results and discussion


Figs. 3-6 show the results from the sensitivity
test on temperature (RMSE), water content (RMSE
and r*) and evaporation
losses (RMSE), respec-

G. Alvetis, P-E. Jansson/Agricultural

52

and Forest Meteorology 88 (1997) 47-56

6
3

3cm

0.6

lcm
7
2

Scm

i
5

10
3

IO

20 cm

cm

cm

20 cm

-10
0
-10

-4

k,

(Iog2mm,d:y)

-4

km

Fig. 3. Root mean square errors of soil temperature at different


depths versus minimum conductivity,
k,, and at different diffusion rates (different d,, values). The curves represent different
values of the factor for the surface humidity correction,
I,$,:
0 = solid line, 0.4 = broken line, 0.8 = dotted line, 1.2 = line with
crosses, 1.6 = line with squares, and 2.0 = line with triangles. For
detailed parameter descriptions, see Eqs. (U-(16).

3cm

-4

-4

(log mm/day)

Fig. 5. Coefficients of determination


for soil water content
different depths (explanations are the same as for Fig. 3).

at

tively. Results from 30-cm depth were quite similar


to those from the 20-cm level and are therefore not
presented here.
Because of the greater simplicity and accuracy in
the temperature measurements compared with the
water content measurements, most of the stress was
laid on the interpretations of the temperatures. However, water content changes recalculated to soil water
losses to the atmosphere were also regarded as useful. Being calculated from changes in water content,
they depend on the accumulated sum of changes and
not on the absolute values of water content. Thereby,

Scm

Root

10 cm

mean square

errors

(mm)

1.0
6

20 cm

0
-10

0
-10

4
k,

2
(log

-4

Fig. 4. Root mean square errors of soil water content at different


depths (explanations are the same as for Fig. 3).

-4
km

mm/day)

(kw

mm/&y)

Fig. 6. Root mean square errors for calculated soil water losses
and simulated evaporation (explanations are the same as for Fig.
3).

G. Alveniis, P-E. Jansson/Agricultural

they might filter out systematic errors connected to


the water content measurements.
3.1. Temperature
The results indicated that a larger simulated vapour
flow, obtained by increasing the diffusion parameter,
dvbr improved the 1e:velof agreement between measured and simulated temperatures (Fig. 3). The best
fit was obtained when using 0.7 and 1.0 for the d,,
parameter. Both values (0.7 and 1.0) are within the
possible range of enhancement factors for sandy
loam as presented in the extensive overview by ten
Berge (1986). However, together with k, values
lower than 10e5 mm/day, a d,, of 1.0 was slightly
better than 0.7, which is close to the value commonly found in the literature. This observation supports theories of vapour diffusion enhancement as
proposed by Philip and de Vries (1957). A test with
even higher vapour diffusion (dvb = 1.5) did, however, worsen the fit between measured and simulated
temperatures, especially in the uppermost 10 cm.
It was interesting to note that reasonable agreement was also obtained in the low-diffusion altematives Cd, = 0.1 and 0.4) when the simulations were
made without correction of the surface vapour pressure (I,$ = 0) at low k, levels. However, a parameter combination resulting in almost non-existing
vapour diffusion can hardly be relevant according to
the literature. Accordingly, as pointed out by Lee
and Pielke (1992) among others, uncorrected vapour
pressure at the surface, i.e., Philips equation uncorrected, should not be used for a drying soil where the
vapour pressure may not be in equilibrium with the
liquid water in the pore system. When a high diffusion rate was allowed (dvb = 1.0) the simulation
without vapour pressure correction resulted in worse
agreement between measurements and simulations
compared with the simulations that account for the
correction term.
Low values of the (Gs parameter, which governs
the near-surface moisture gradient, demonstrated a
high sensitivity to the minimum hydraulic conductivity, k,, whereas high values rendered the k, parameter unimportant. The pronounced near-surface moisture gradient, created by the use of a high I& value,
reduced evaporation losses markedly in spite of the
high hydraulic conductivity. The interplay between

and Forest Meteorology 88 (1997) 47-56

53

the 4 and k, parameters implied that a good


agreement between measured and simulated temperatures could be obtained from a number of different
combinations of the two parameters. However, at
high diffusion rates (dvb = 1.01, which were regarded as reasonable, the root mean square errors
indicated that k, values < 10m4 mm/day and lclg
values higher than 0.4 were most appropriate.
The calculated coefficients of determination, r2,
for temperature supported the interpretations based
on the RMSE curves.
3.2. Water content
The water content simulations were more tricky to
interpret compared with the temperature simulations.
In addition, any systematic errors connected with the
TDR measurements could have influenced the RMSE
calculations. One possible error could be connected
with the 10% clay content of the studied soil. Dirksen and Dasberg (1993) showed that the most commonly used TDR calibration equation (Topp et al.,
19801, which also was used in this study, may cause
underestimations of the water contents in fine-textured soils. Thus, to avoid drawing any erroneous
conclusions the coefficients of determination were
also considered.
At 3- and 7-cm depth the water contents, simulated without the vapour pressure correction ( I)~= 01,
generally deviated strongly from measured water
contents (Fig. 4). Simulated water contents decreased
faster than measured ones owing to comparably high
evaporation losses caused by overestimated surface
vapour pressures. However, with increasing I+$ values, i.e., an increasing near-surface gradient, the
agreement between measured and simulated water
contents was improved. The lower surface vapour
pressure, induced by a higher I,$ value, resulted in
reduced evaporation losses from the soil. When intermediate or high diffusion rates were presumed, a
I& value above 0.8 had to be used to prevent soil
water losses to the atmosphere from being too high.
Using a minimum hydraulic conductivity (k,)
below lo- mm/day at lo- and 20-cm depths, the
level of agreement between measurements and simulations increased with increasing vapour diffusion
(up to d, = 0.7-1.01, as was also true for the
temperature. Exceptions were found at 3- and 5-cm

54

G. Alum&,

P-E. Jansson / Agricultural and Forest Meteorology 88 (1997) 47-56

depths, where root mean square values increased


with higher diffusion rates. At 3-cm depth, however,
this trend was contradicted by the coefficients of
determination
(Fig. 51, which expressed a better
agreement between measured and simulated values
at d,, values above 0.7 than at lower values of the
parameter. However, no conclusions supporting the
theories of vapour flow enhancement could be drawn
from these results. As for the temperature, a further
increase in vapour diffusion (dvb = 1.5) resulted in
poor agreement between measured and simulated
values.
The root mean squares indicated that the simulated water contents were highly sensitive to the
level of the minimum hydraulic conductivity,
k,,
within a certain intermediate
range. Below 10V6
mm/day, however, a further decrease in k, did not
affect the soil water simulation markedly. This could
be due either to the fact that the unsaturated conductivity of the soil never decreased to a lower level,
i.e., the k, was not reached, or to the water transport being so small at a conductivity of lop6 mm/
day that further decreases were insignificant.
The
coefficient of determination was in accordance with
the root mean squares, showing no particular changes
below k, = 10e6 mm/day.
The obvious interplay between the water vapour
flow and the liquid water flow was somewhat complicated. A high value of the diffusion coefficient in
combination with a low k, resulted in the uppermost
levels drying too fast. The high diffusion rate caused
high evaporation losses, while the water supply from
deeper layers was limited by the low conductivity.
With increasing minimum conductivities
the water
supply from the subsoil increased. At a k, of 10e310e2 mm/day
the supply and losses had reached
rates that resulted in a considerably better agreement
between measured and simulated water contents.
With a further increased k,, however, the soil water
content at the 3- and 7-cm depths were overestimated. The coefficients of determination,
r2, had
distinct maximas, indicating good agreements, around
a minimum conductivity of low3 mm/day, which is
in accordance with what was indicated by the root
mean squares. However possible minimum conductivities as high as 1O-3-1O-2
mm/day
are not in
line with the reasonable range indicated by the temperature comparisons. Besides, referring to measure-

ment uncertainties mentioned earlier and to the uncertainties in the k(JI) relation as discussed below,
the k, interpretation is afflicted with uncertainties.
Above lo- mm/day
the coefficient of determination dropped dramatically to approach 0 with further
increased
k, values irrespective of the diffusion
rate. This indicates that such high minimum conductivities could not be used.
When evaluating the k, parameter it should be
kept in mind that the relation between unsaturated
conductivity and soil water content in the dry range
is uncertain. Because direct measurements
are time
consuming, expensive and difficult to perform (e.g.,
Ragab et al., 1981, or van Genuchten et al., 1989)
the k( I/J) relation in this case was calculated from
measured saturated conductivity, k,, using the equation of Mualem (1976). However, as pointed out by
Khaleel et al. (1995) among others, using k, as a
single matchpoint for the k($) curve may result in
inadequate characterization of the relationship in the
dry moisture range. Although it is true that the k(JI)
curve was also compared with conductivities calculated from water contents and tensions measured
during a drying period (Fig. lb), the scatter in the
measured-calculated
values and the narrow range
within which the tensiometers can be used leaves
room for different interpretations
of the k( I,!I) relation. Slightly changing the shape of the kc+) function might give a result similar to that obtained by
changing the k, value. Methods using soil evaporation as a boundary condition for calculating k( I/I> in
the dry range, e.g., the hot-air method (Arya et al.,
19751, naturally cannot be used as validation since
they include an unknown contribution
of vapour
flow.
3.3, Soil evaporation
As for the water contents, no clear indications of
enhanced vapour diffusion were given by comparisons between measured soil water losses at O-30 cm
depth and simulated soil evaporation (Fig. 6). At
high diffusion rates (dvb = 1.0) the vapour flow to
the surface caused an overestimation
of the simulated evaporation
irrespective of the level of the
lowest possible hydraulic conductivity (k, >. Similarly, low diffusion rates (dvb = 0.4 or 0.1) resulted
in underestimated
simulated evaporation owing to

G. Alueniis, P-E. Jansson/Agricultural

and Forest Meteorology

88 (1997) 47-56

55

the limited amounts of moisture that could be transported to the surface. With increasing minimum
hydraulic conductivity ( k, >, however, the moisture
deficiency at the surface turned into a surplus, leading to overestimated evaporation, particularly when
small moisture gradients close to the surface (low I,$
values) were presumed to exist.
The dependence of simulated evaporation on the
slope of the near-surface moisture gradient was
clearly demonstrated in this study. A good agreement
between simulated e,vaporation and water losses calculated from measured soil water contents could
only be obtained when a I,$ value of 1.2 or higher
was used. This result restricts the possible range of
I,$ values more strongly than restrictions obtained
with the temperature and water content variables,
thus implying that rsurface vapour pressure already
starts to decrease at surface moisture potentials as
high as - 100 hPa.
As for the temperatures, minimum conductivity
levels above 10-5-10-4 mm/day resulted in poor
agreement between simulated and measured water
losses.

and evaporation simulations did not confirm their


ideas.
Water and vapour flows act together and are hard
to distinguish between. Close to the surface a low k,
can be compensated by a high d,, and vice versa.
However, in this soil, too low conductivities in combination with high diffusion rates may lead to an
unrealistically steep transition from dry to wet conditions in the topsoil.
To investigate the generality of the model approach, we plan to carry out additional tests on other
moisture conditions and soil types.

4. Conclusions

Alvenls, G., Johnsson, H., Jansson, P.-E., 1986. Meteorological


conditions and soil climate of four cropping systems. Measurements and simulations from the project Ecology of Arable
Land. Department of Ecology and Environmental
Research,
Report 24. Swedish University of Agricultural Sciences, Uppsala, Sweden.
Arya, L.M., Farrell, D.A., Balke, G.R., 1975. A field study of soil
water depletion patterns in presence of growing soybean roots:
1. Determination of hydraulic properties of the soil. Soil. Sci.
Sot. Am. Proc. 39,424-430.
Barton, LJ., 1979. A parameterization
of the evaporation from
nonsaturated surfaces. J. Appl. Meteor. 18, 43-47.
Camillo, P.J., Gurney, R.J., Schmugge, T.J., 1983. A soil and
atmospheric boundary layer model for evapotranspiration
and
soil moisture studies. Water Resource Res. 19, 371-380.
Cary, J.W., 1963. Onsagers relation and the non-isothermal diffusion of water vapor. J. Phys. Chem. 67, 126-129.
Cass, A., Campbell, G.S., Jones, T.L., 1984. Enhancement
of
thermal water vapor diffusion in soil. Soil Sci. Sot. Am. J. 48,
25-32.
Deardorff, J.W., 1978. Efficient prediction of ground surface
temperature and moisture, with inclusion of a layer of vegetation. I. Geophys. Res. 83, 1889-1903.
DekiC, L., MihailoviC, D.T., Rajkovic, B., 1995. A study of me
sensitivity of bare soil evaporation schemes to soil surface
wetness, using the coupled soil moisture and surface temperature prediction model BARESOIL. Meteorol. Atmos. Phys.
55, 101-112.

Due to the higher accuracy of temperature measurements compared with soil water measurements,
soil temperature was more useful than water content
for evaluating methods of estimating soil surface
energy balance. However, evaporation calculated
from changes in the soil water storage was a useful
complement that could be used to restrict the range
of some crucial and uncertain properties of the system.
Soil temperature, soil water content and evaporation all indicated that the use of the correction factor,
s, is necessary to prevent soil evaporation rates
CF,
from being too high. +s values as high as 1.2 had to
be used for the soil in this study. This level is in
agreement with the results presented by Schelde et
al. (1997).
Good agreements between measured and simulated temperatures were achieved when d, values
close to 1.0 were used. This result supports the
theories of vapour flow enhancement proposed by
Philip and de Vries. However, the soil water content

Acknowledgements

We would like to thank Lave Persson and Gijsta


Ljung for valuable help with the experiment set up
and Christina Ghman for carefully performed soil
analyses. The study was funded by the Swedish
Natural Science Research Council.

References

56

G. Alveniis, P-E. Jansson /Agricultural

Dirksen, C., Dasberg, S., 1993. Improved calibration of time


domain reflectometry soil water content measurements. Soil
Sci. Sot. Am. J. 57, 660-667.
Jansson, P.-E., 1991. Simulation model for soil water and heat
conditions. Description of the SOIL model. Rep. 165, Department of Soil Sciences, Swedish University of Agricultural
Sciences, Uppsala, Sweden.
Khaleel, R., Relyea, J.F., Conca, J.L., 1995. Evaluation of van
Genuchten-Mualem relationships to estimate unsaturated hydraulic conductivity at low water content. Water Resource
Res. 31, 2659-2668.
Kondo, J., Saigusa, N., Sato, T., 1990. A parameterization of
evaporation from bare soil surfaces. J. Appl. Meteor. 29,
383-387.
Kondo, J., Saigusa, N., Sato, T., 1992. A model and experimental
study of evaporation from bare-soil surfaces. J. Appl. Meteor.
31, 304-312.
Lee, T.J., Pielke, R.A., 1992. Estimating the soil surface specific
humidity. J. Appl. Meteor. 31, 480-484.
Mafouf, J.F., Noilhan, J., 1991. Comparative study of various
formulations of evaporation from bare soil using in situ data.
J. Appl. Meteor. 30, 1354-1365.
McCumber, MC., Pielke, R.A., 1981. Simulation of the effects of
surface fluxes of heat and moisture in a mesoscale numerical
model: 1. Soil layer. J. Geophys. Res. 86, 9929-9938.
MihaloviC, D.T., Pielke, R.A., RajkoviC, B., Lee, T.J., Jeftic, M.,
1993. A resistance representation of schemes for evaporation
from bare and partly plant-covered surfaces for use in atmospheric models. J. Appl. Meteor. 32, 1038-1054.
Mualem, Y., 1976. A new model for predicting the hydraulic
conductivity of unsaturated porous media. Water Resource
Res. 12, 513-522.

and Forest Meteorology

88 (1997) 47-56

Nappo, C.J., 1975. Parameterization of surface moisture and


evaporation rate in a planetary boundary layer model. J. Appl.
Meteor. 14, 289-296.
Philip, J.R., 1957. Evaporation, and moisture and heat fields in the
soil. J. Meteor. 14, 354-366.
Philip, J.R., de Vries, D.A., 1957. Moisture movement in porous
materials under temperature gradients. Trans. Am. Geophys.
Union 3, 222-232.
Ragab, R., Feyen, J., Hillel, D., 1981. Comparative study of
numerical and laboratory methods for determining the hydraulic conductivity function of a sand. Soil Sci. 131,375-388.
Richards, L.A., 1931. Capillary conduction of liquids through
porous media. Physics 1, 318-333.
Schelde, K., Thomsen, A., Heidmann, T., Schjotig,
P., 1997.
Diurnal fluctuations of water and heat flows in a bare soil. (in
press).
ten Berge, H.F.M., 1986. Heat and water transfer at the bare soil
surface. Aspects affecting thermal imagery. PhD thesis, Department of Soil Science and Plant Nutrition, Wageningen
Agricultural University, The Netherlands.
Topp, G.C., Davis, J.L., Annan, A.P., 1980. Electromagnetic
determination of soil water content: Measurements in coaxial
transmission lines. Water Resource Res. 16, 574-582.
van Genuchten, M.Th., Kaveh, F., Russell, W.B., Yates, S.R.,
1989. Direct and indirect methods for estimating the hydraulic
properties of unsaturated soils. In: Bouma, J., Bregt, A.K.
(Eds.), Land qualities in space and time. Proceed. Symps. by
ISSS at Wageningen, Netherlands. Wageningen, Netherlands,
Pudoc, pp. 61-72.
Yasuda, N., Toya, T., 1981. Evaporation from non-saturated
surface and surface moisture availability. Pap. Meteor. Geophys. 32, 89-98.

Das könnte Ihnen auch gefallen