Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Agricultural Water Management 96 (2009) 1587–1595

Contents lists available at ScienceDirect

Agricultural Water Management


journal homepage: www.elsevier.com/locate/agwat

Simulation of soil water dynamics under subsurface drip irrigation from


line sources
S. Elmaloglou *, E. Diamantopoulos
Department of Natural Resources Management & Agricultural Engineering, Agricultural University of Athens, 11855 Athens, Greece

A R T I C L E I N F O A B S T R A C T

Article history: A mathematical model which describes water flow under subsurface drip lines taking into account root
Received 28 November 2008 water uptake, evaporation of soil water from the soil surface and hysteresis in the soil water
Accepted 12 June 2009 characteristic curve u(H) is presented. The model performance in simulating soil water dynamics was
Available online 17 July 2009
evaluated by comparing the predicted soil water content values with both those of Hydrus 2D model and
those of an analytical solution for a buried single strip source. Soil water distribution patterns for three
Keywords: soils (loamy sand, silt, silty clay loam) and two discharge rates (2 and 4 l m1 h1) at four different times
Subsurface drip irrigation modelling
are presented. The numerical results showed that the soil wetting pattern mainly depends on soil
Discharge rate
Soil hydraulic properties
hydraulic properties; that at a time equal to irrigation duration decreasing the discharge rate of the line
Deep percolation sources but maintaining the applied irrigation depth, the vertical and horizontal components of the
Evapotranspiration wetting front were increased; that at a time equal to the total simulation time the discharge rate has no
Hysteresis effect on the actual transpiration and actual soil evaporation and a small effect on deep percolation. Also
Soil water distribution patterns the numerical results showed that when the soil evaporation is neglected the soil water is more easily
taken up by the plant roots.
ß 2009 Elsevier B.V. All rights reserved.

1. Introduction a two- and three-dimensional unsaturated water flow from buried


point sources and spherical cavities. Warrick et al. (1980) reported
Subsurface drip irrigation (SDI) under certain conditions is the a mathematical model to describe three-dimensional linearized
most efficient method of irrigating agricultural crops and land- moisture flow with root extraction under steady conditions for
scape. A comprehensive review of published information on various subsurface sources. Ben-Asher and Phene (1993) presented
subsurface drip irrigation is given by Camp (1998) to determine the a numerical model to analyze two-dimensional water flow for
state of the art on the subject. SDI systems are used to provide surface and subsurface drip systems. Or (1996) used a stochastic
water to plant roots while maintaining a relatively dry soil surface, approach to develop analytical expressions relating variations in
which ensures that the applied water becomes available to a soil hydraulic properties to expected variability in matric potential
substantial fraction of the plant root system. With SDI the farmers and relative saturation, which could be used to determine the
can achieve more efficient water use, less water quality hazards, number and placement of tensiometers and Time Domain
improved opportunities for use of degraded waters and greater Reflectometry (TDR) probes. Coelho and Or (1996) presented a
water application uniformity (Lamm and Camp, 2007). Also SDI parametric model for two-dimensional flow and uptake by corn for
improves the plants’ health, farming operations and management. four plant-emitter configurations.
Drip irrigation using buried emitters has the potential to save One of the important aspects of planning and managing a SDI
irrigation water by reducing soil surface wetting and thus reducing system is the soil moisture movement pattern below it. The depth
evaporation. Evett et al. (1995) hypothesized that improved yields of the lateral below soil surface, emitter spacing and system
from SDI systems are most likely due to more water being available pressure are of utmost importance for delivering the required
to the plants, as compared to a surface drip system because of less amount of water to the plant. Wetting pattern can be obtained by
evaporation in a subsurface drip system. either direct measurement of soil wetting in the field, which is site-
In general, water flow under drip irrigation is a three- specific, or by simulation using some models (Ismail et al., 2006;
dimensional problem but becomes two-dimensional when used Bhatnagar and Chauhan, 2008). Recently, Cote et al. (2003), Skaggs
as a line source. Philip (1968) developed a mathematical theory for et al. (2004), and Gardenas et al. (2005) used the Hydrus 2D model
(Simunek et al., 1999) in order to simulate soil water dynamics
under subsurface drip irrigation.
* Corresponding author. Fax: +30 2105294081. In this paper a mathematical model which describes water flow
E-mail address: elma@aua.gr (S. Elmaloglou). under subsurface drip lines taking into account root water uptake,

0378-3774/$ – see front matter ß 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.agwat.2009.06.010
1588 S. Elmaloglou, E. Diamantopoulos / Agricultural Water Management 96 (2009) 1587–1595

evaporation of soil water from the soil surface and hysteresis in the corresponding to initial volumetric water content ui uniform in the
soil water characteristic curve u(H) is presented. The model horizontal and vertical direction:
performance in simulating soil water was evaluated by comparing
the predicted soil water content values with both those of the uðx; z; t ¼ 0Þ ¼ ui ; Hðx; z; t ¼ 0Þ ¼ Hi (5)
Hydrus 2D model and those of the analytical solution of Warrick The condition in the lower right margin of the flow region is
and Lomen (1976) for a single buried strip source. defined in such a way that the discriminative region be limited to
the soil region where the variable U changes with time. The
2. Methodology discriminative region increases with time as long as the wetting
front expands, in such a way that it is restricted to the area ABCD
2.1. The physical model (Fig. 1).
Condition at the moving boundary G3 (Fig. 1) of flow region:
Fig. 1 illustrates the three-dimensional physical model. Drip
irrigation is applied 20 cm below the soil surface, from a horizontal For t > 0; U ¼ U i ; at z ¼ ZðtÞ; 0 < x < LS=2 (6)
line source, of width 2x0. The plane flow symmetry allows the
where Z(t) is the coordinate of the moving boundary and LS the
physical model to be examined in one of the infinite vertical planes
lateral spacing.
perpendicular to the length of the line source and determined by
Conditions at constant boundaries of flow region:
the x and z axes.

(a) In the plane of symmetry G2 (x = LS/2), there is a no-flow


2.2. The mathematical model
condition:
The water flow equation that describes the physical model is: @U
For t > 0; ¼ 0; at 0  z  ZðtÞ
qx ¼  (7)
@x
@U @2 U @2 U @U
FðUÞ ¼ þ  GðUÞ  SðH; zÞ (1) (b) In the plane of symmetry G1 (x = 0), there is a no-flow
@t @z2 @x2 @z
condition:
where the coefficients F(U) and G(U) are given by: @U
For t > 0; qx ¼  ¼ 0; at 0  z  ZðtÞ (8)
1 du 1 dK @x
FðUÞ ¼ FðHÞ ¼ and GðUÞ ¼ GðHÞ ¼ (2)
KðHÞ dH KðHÞ dH (c) At the soil surface (z = 0), the condition without rainfall is:

and Rubin (1968): @U


For t > 0; qz ¼  þ K ¼ Ea ; at 0 < x  LS=2 (9)
Z H @z
UðHÞ ¼ KðHÞdH (3) where Ea is the actual evaporation rate.
Hi

and x, z are the Cartesian coordinates (L) with z positive The subsurface drip irrigation from line sources was simulated
downwards, H is the pressure head (L), K is the hydraulic by considering the irrigation lines placed at the plane of symmetry
conductivity (L/T), u is the volumetric water content of the soil G1, 20 cm below the soil surface, so that half of the water
(L3/L3), t is time since the beginning of flow (T), Hi is the initial application rate was considered to enter in the simulated area
pressure head (L), S is a distributed sink function representing the ABCD. The irrigation flux was included as a source term at the node
water uptake by the roots (L3 L3 T1). where the line is placed.

2.3. Initial and boundary conditions


2.4. Implementation
The initial condition is:
For the numerical solution of the soil water flow equation
Uðx; z; t ¼ 0Þ ¼ U i (4) [Eq. (1)], the Alternating Direction Implicit (ADI) method was used,
which requires the completion of two time stages for a total
applying cycle. The simulated region is divided into a network of
equally spaced grids (Dx = Dz = 2 cm), where each point is
designated by the subscripts i and j, referring to the horizontal
and vertical direction respectively, as illustrated in Fig. 2. The
symmetry axis BD is inserted between the columns (i = NR, j) and
(i = NR + 1, j). The moving boundary G3 is inserted between the
lines (i, j = M) and (i, j = M + 1). The bottom limit of the rooting zone
is inserted in the line (i, j = MR).
The space–time continuum (x, z, t) is covered by grid points (xi,
zi, t‘) (Fig. 2), where:

xi ¼ ði  1:5ÞDx i ¼ 2; 3; . . . ; NR þ 1
z j ¼ ð j  2ÞDz j ¼ 2; 3; . . . ; M þ 1
X ‘ (10)
t‘ ¼ Dt ‘ with t 0 ¼ 0
‘¼1

Here, the superscript ‘ is used to specify the time level and Dt‘ is
the time step at the ‘th level of time. In the ADI approach, the
solution for the next time level (t‘+1) was obtained from the known
values at the time level (t‘) in two stages. During the first stage, the
Fig. 1. The three-dimensional physical model. solution was allowed to advance from t‘ to t‘+1/2 with implicit
S. Elmaloglou, E. Diamantopoulos / Agricultural Water Management 96 (2009) 1587–1595 1589

simplification made by Kool and Parker (1987) for the empirical


model of Scott et al. (1983). Elmaloglou and Diamantopoulos
(2008) explain this method in detail.
Three of the twelve USDA soil classes (Schaap and Leij, 1998)
covering a wide range of cultivated soils were selected and the
class average values of the u(H) and K(H) relationships according to
Van Genuchten (1980) were used. The analytical expressions of
those relationships are:
8
< us  ur
ur þ m H<0
uðHÞ ¼
: ð1 þ jaHjn Þ (13)
us ; H  0

m 2
KðHÞ ¼ K s Se0:5 ½1  ð1  Sen=ðn1Þ Þ  (14)

uðHÞ  ur 1
where Se ¼ ; m¼1 (15)
us  ur n

Fig. 2. Schematic representation of the computing domain.


and ur is the residual water content (L3/L3); us is the saturated water
content (L3/L3); Ks is the saturated hydraulic conductivity (L/T); Se
is the effective saturation (dimensionless) and n, a are shape
differences used in the z-direction and explicit differences in the x- parameters (a(L1)).
direction, while in the second stage the solution advances from t‘+1/ The selected soil classes, from the more coarse-grained to more
2
to t‘+1 with implicit differences in the z-direction and explicit fine-grained, were the loamy sand, silt and silty clay loam. To
differences in the x-direction. determine the main drying curve from the main wetting curve, the
The governing Eq. (1) is discretized for the first stage as follows: empirical model introduced by Scott et al. (1983) and modified by
‘þ1=2 Kool and Parker (1987) was used. We assumed that ad ¼ aw =2 with
Ui; j  Ui;‘ j ‘
Uiþ1; ‘ ‘
j  2Ui; j þ Ui1; j aw and ad the a parameter in the relationship of Van Genuchten
Fi;‘ j ¼
Dt ‘=2 Dx 2 (1980) for the main wetting and the main drying curve. For the
‘þ1=2 ‘þ1=2 ‘þ1=2
Ui; jþ1  2Ui; j þ Ui; j1 K(H) relation the chosen parameter vector was (u s ; ur ; aw ; n; K s ).
þ The values of the parameters included in the van Genuchten
Dz2
‘þ1=2 ‘þ1=2 equations are summarized in Table 1. The values of initial
Ui; jþ1  Ui; j1
 G‘i; j  S‘i; j þ W i; j (11) volumetric water content ui were 0.164 cm3/cm3 for the loamy
2Dz sand, 0.198 cm3/cm3 for the silt and 0.222 cm3/cm3 for the silty
and for the second stage as: clay loam. The above values were chosen so that at time t = 0, all
soils had the same value of effective saturation (Se).

‘þ1=2 2.6. Validation of the mathematical model


‘þ1=2
Ui;‘þ1
j
 Ui; j ‘þ1
Uiþ1; j
 2Ui;‘þ1
j
‘þ1
þ Ui1; j
Fi; j ¼
Dt ‘=2 Dx2
‘þ1=2 ‘þ1=2 ‘þ1=2 Due to lack of an analytical solution for the differential
Ui; jþ1  2Ui; j þ Ui; j1
þ flow Eq. (1) subject to the conditions formulated above, one
Dz2 way to control the convergence of its numerical solution is
‘þ1=2 ‘þ1=2
U
‘þ1=2 i; jþ1
 Ui; j1 ‘þ1=2 to compare the solution of the linearized form of Eq. (1) with
 Gi; j  Si; j þ W i; j
2Dz an existing analytical solution of a closely related simpler problem.
(12) In that comparison, evapotranspiration is disregarded and the
characteristic curves [u(H), K(H)] are assumed to be given by:
As shown in Fig. 2 the line source is placed at the node (2, jsi)
perpendicular to the plane of the simulated area. The irrigation flux ½KðHÞ  K s þ us k
uðHÞ ¼ ; KðHÞ ¼ K s expðaHÞ (16)
was included as a source term (Wi,j) in the second part of Eqs. (11) k
and (12) at the node (2, jsi) where the line source was placed. W(2,
where k = dK/du and a (1/L) are constants. The flow equation then
jsi) is the volumetric irrigation flux per unit volume of soil (T1).
reduces to the linear equation of Warrick and Lomen (1976):
2.5. Hysteresis and soil physical characteristics @f k 2 @f
¼ r fk (17)
@t a @z
The phenomenon of hysteresis in the soil water characteristic RH
curve was incorporated in the mathematical model with the where f ¼ 1 KðHÞdH ¼ Ka :

Table 1
Values of the shape parameter (n), the saturated (us), residual (ur) water contents, the saturated hydraulic conductivity (Ks) and the scaling factors (aw ) and (ad) for the main
wetting and drying curve, respectively (Elmaloglou and Diamantopoulos, 2008).

Soil aw (1/cm) ad (1/cm) n () us (cm3/cm3) ur (cm3/cm3) Ks (cm/h)

Loamy sand 0.03467 0.017335 1.7378 0.390 0.049 4.383


Silt 0.00661 0.003305 1.6596 0.489 0.05 1.819
Silty clay loam 0.00832 0.00416 1.5136 0.482 0.09 0.467
1590 S. Elmaloglou, E. Diamantopoulos / Agricultural Water Management 96 (2009) 1587–1595

Table 2
Values of the parameters of Eq. (16) and textural values for the Sable DEK soil.

Soil a (1/cm) k (cm/h) us (cm3/cm3) Ks (cm/h) Fine particles percentage (0–50 mm) Argile percentage (0–2 mm)

Sable DEK 0.073 33.42 0.3 5.85 18 8

The numerical solution was compared with the analytical numerical solution of the linearized flow equation agrees with the
solution of Warrick and Lomen (1976) for a subsurface strip analytical solution of Warrick and Lomen (1976).
source: The mathematical model was also compared with the results of
pffiffiffiffi Z ! the Hydrus 2D model, for a subsurface line source. For this
p exp Z T 1=2 Z2 comparison the loamy sand soil was used. Hydrus 2D uses the
F2B ðX; Z; TÞ ¼ t exp t 
4X 0 0 4t Galerkin finite-element method to solve the Richards equation.
    
ðX þ X 0 Þ ðX  X 0 Þ Simunek et al. (1999) explain the solution in detail. The comparison
 er f pffiffiffi  er f pffiffiffi dt (18) was made because Hydrus 2D is a widely used numerical model and
2 t 2 t
it has been tested under various experiments including also
where F, X, X0, Z, T are the dimensionless coordinates of f, x, x0, z, t subsurface irrigation (Skaggs et al., 2004; Gardenas et al., 2005).
respectively, calculated from the following relationships: Fig. 4 shows soil water profiles for two different times, t = 4.5 h and
8pf ax ax0 az akt t = 9 h, for two different distances, x = 1 cm and x = 15 cm from the
F¼ ; X¼ ; X0 ¼ ; Z¼ ; T¼
plane of symmetry (x = 0). The discharge rate was again 2 l m1 h1
Q 2 2 2 4
and the agreement between the Hydrus 2D model and the model
To compare the numerical solution with the analytical solution, presented here is very good for both times and distances.
the Sable DEK soil (Ababou, 1981) was used. Sable DEK is a fine At last the validity of the presented numerical solution can be
permeable sand containing a non-negligible percentage of fine controlled by the criterion of maintenance of volume (Elma-
particles (Table 2). The wetting front development for a discharge loglou and Diamantopoulos, 2008). In the numerical simulations
rate of 2 l m1 h1 is presented in Fig. 3. It can be seen that the presented here the relative volume balance error remained
under 1%.

2.7. Evapotranspiration and simulation inputs

The mathematical model considers the actual evaporation rate


Ea from the soil surface as a declining exponential function with

Fig. 4. Water content profiles for two horizontal distances (x = 1 cm and x = 15 cm)
Fig. 3. Wetting front development as a function of infiltration time for Sable DEK. from the plane of symmetry (x = 0) and for two times (a) t = 4.5 h, (b) t = 9 h.
S. Elmaloglou, E. Diamantopoulos / Agricultural Water Management 96 (2009) 1587–1595 1591

Fig. 5. Soil water distribution patterns for loamy sand soil, with discharge rate of 2 l m1 h1, for four different times (a) ti, (b) 24 h, (c) 48 h, and (d) tend = 72.4 h.

decreasing pressure head H: where a(H) is a dimensionless reduction factor related to the effect
of water stress given by Feddes et al. (1978).
Ea ¼ Ep edH (19)
aðHÞ ¼ 0 for H  H1 or H  H4 (21)
where d is a dimensionless empirical constant, Ep is the potential
evaporation rate of the soil surface being 0.01 cm/h during a 12 h H  H1
daytime and 0 cm/h during the next 12 h nighttime. In this way the aðHÞ ¼ for H2 < H < H1 (22)
H2  H1
daily potential soil evaporation (PE) was calculated equal to 0.12 cm.
The transpiration rate is expressed by the distributed sink term,
aðHÞ ¼ 1 for H3  H  H2 (23)
S, which is a function of maximum water extraction rate Smax at
depth z and the water pressure head H (Feddes et al., 1978): H  H4
aðHÞ ¼ for H4 < H < H3 (24)
SðH; zÞ ¼ aðHÞSmax ðzÞ (20) H3  H4

Fig. 6. Soil water distribution patterns for loamy sand soil, with discharge rate of 4 l m1 h1, for four different times (a) ti, (b) 24 h, (c) 48 h, and (d) tend = 72.4 h.
1592 S. Elmaloglou, E. Diamantopoulos / Agricultural Water Management 96 (2009) 1587–1595

Fig. 7. Soil water distribution patterns for silt soil, with discharge rate of 2 l m1 h1, for four different times (a) ti, (b) 24 h, (c) 48 h, and (d) tend = 78.0 h.

where H1 = 10 cm is the pressure head below which the roots start where a is the maximum extraction rate at the surface, b is the
to extract water from the soil, H2 = 25 cm is the pressure head reduction coefficient. The values of a and b were set to
below which the roots start to extract water optimally from the soil, 0.15  102 cm3 cm3 h1 and 0.22  104 cm1 h1 respectively,
H3 = 400 cm is the pressure head below which the roots cannot while g was set to 1.00 during the day and 0.20 during the night.
optimally extract water from the soil, and H4 = 15,000 cm is the The root depth was taken as 60 cm. According to the values of the
pressure head below which the roots cannot extract soil water. parameters (a, b, g) in Eq. (25) and the root depth, the calculated
The relationship of Smax to the depth, z, assumes a linear daily potential transpiration (PT) was 0.72 cm. From the above
variation given by the formula of Hoogland et al. (1981), combined paragraphs, the calculated potential evapotranspiration (PET) was
with a coefficient g, which represents the daily variation of the 0.84 cm which is a common value for semi-arid regions. The value
potential transpiration: of Ep = 0.01 cm/h was chosen so that the ratio PE/PET has a value
around 14% which corresponds to real conditions. The value of the
Smax ðzÞ ¼ g ða  bjzjÞ (25) constant d of Eq. (19) was chosen equal to 104 for the loamy sand,

Fig. 8. Soil water distribution patterns for silt soil, with discharge rate of 4 l m1 h1, for four different times (a) ti, (b) 24 h, (c) 48 h, and (d) tend = 78.0 h.
S. Elmaloglou, E. Diamantopoulos / Agricultural Water Management 96 (2009) 1587–1595 1593

Fig. 9. Soil water distribution patterns for silty clay loam soil, with discharge rate of 2 l m1 h1, for four different times (a) ti, (b) 24 h, (c) 48 h, and (d) tend = 84.0 h.

8  105 for the silt and 2.5  105 for the silty clay loam so that examined. The total amount of irrigation water is 18 l/m.
the ratio of actual evaporation to actual evapotranspiration, as it is Figs. 5–10 present the variation in soil water distribution patterns
calculated from the numerical results, for the three soils was 14% at at four different times [irrigation duration ti (figures a), 24 h
a time equal to the total simulation time. The total simulation time (figures b), 48 h (figures c) and total simulation time tend (figures
was defined as the time needed for the mean water content in the d)], for the three soils and the two discharge rates Q = 2 l m1 h1
effective root zone to reach the initial mean water content. and Q = 4 l m1 h1. The irrigation duration ti varies according to
discharge rate (Table 3).
3. Results and discussion From Figs. 5a, 6a, 7a, 8a, 9a and 10a it can be seen that at time
t = ti the smaller discharge rate (Figs. 5a, 7a and 9a) gives larger soil
Soil wetting pattern around a buried line source of water water movement in both vertical and horizontal direction for all
application is mainly dependent on soil hydraulic properties, soils. This could be attributed to the limitation of the soil
discharge rate of laterals, duration of water application, depth of infiltration rate. The drip line with the higher discharge rate
placement of laterals and root water uptake. In this work the (4 l m1 h1) delivers the total amount of irrigation water (18 l/m)
effects of soil hydraulic properties and discharge rate were in a short time (4.5 h) and cannot infiltrate fast enough into the

Fig. 10. Soil water distribution patterns for silty clay loam soil, with discharge rate of 4 l m1 h1, for four different times (a) ti, (b) 24 h, (c) 48 h, and (d) tend = 84.0 h.
1594 S. Elmaloglou, E. Diamantopoulos / Agricultural Water Management 96 (2009) 1587–1595

Table 3 Table 5
Combinations of discharge rate and irrigation durations. Deep percolation, actual transpiration and actual evaporation at time t = tend as a
percentage of the applied irrigation depth (potential transpiration is taken equal to
Discharge rate (l m1 h1) Irrigation duration ti (h) Irrigation depth (mm) 0.54 cm).
2 9.0 30 Discharge rate Soil (%)
4 4.5 30
Deep percolation 2 l m1 h1 Loamy sand 19.2
Silt 8.4
Silty clay loam 0.6
4 l m1 h1 Loamy sand 18.8
Silt 8.0
Silty clay loam 0.6

Actual transpiration 2 l m1 h1 Loamy sand 66.0


Silt 75.5
Silty clay loam 82.9
4 l m1 h1 Loamy sand 66.4
Silt 75.8
Silty clay loam 82.9

Actual evaporation 2 l m1 h1 Loamy sand 14.8


Silt 16.1
Silty clay loam 17.7
4 l m1 h1 Loamy sand 14.8
Silt 16.2
Silty clay loam 17.7

throughout the effective root zone increases from the coarser soil
(loamy sand) to the finer soil (silty clay loam).
For the three soils and for both discharge rates (2 and
Fig. 11. The evolution of the average water content inside the root zone for the three 4 l m1 h1), the deep percolation below the root zone, the actual
soils and for the two discharge rates. transpiration and the actual soil evaporation are given in Table 4
at time t = tend as a percentage of the applied irrigation depth.
surrounding soil. Therefore water accumulates around the line From Table 4 it can be seen that, deep percolation decreased from
source. On the other hand, the lower discharge rate delivers the the coarser to the finer soil. On the contrary, actual transpiration
same amount of water in a larger time (9 h) and so allows lateral as and actual evaporation increased from the coarser to the finer soil.
well as vertical distribution of the water resulting in a greater In addition, from Table 4 it can be seen that the discharge rate has
wetted area with lower soil water content. From Figs. 5–10 it can no effect on the actual transpiration and actual evaporation and
be seen that for all times and the three soils a higher soil water only a small effect on deep percolation. Table 5 shows the deep
content around the line source can be observed at a higher percolation, the actual transpiration and the actual soil evapora-
discharge rate. tion percentages when the daily potential transpiration is taken
From Figs. 5–10 it can be seen that for all times and both equal to 0.54 cm (75% of 0.72 cm). The comparison between
discharge rates downward soil water movement is greater than Tables 4 and 5 shows that the decrease of actual transpiration is
upward soil water movement because capillary forces are small almost equally distributed to deep percolation and actual
compared with gravity forces. This is more evident in the coarser evaporation. For example, in the case of the loamy sand and for
soil (loamy sand), because of the greater relative gravitational a discharge rate of 2 l m1 h1, actual transpiration decreased
influence in this soil type. From Fig. 11 it can also be seen that, for from 72.7 to 66.0% while deep percolation increases from 15.4 to
the same initial effective saturation Se, the average water content 19.2% and actual evaporation increases from 11.9 to 14.8%. Table 6
shows the deep percolation and actual transpiration percentages
Table 4 for PE = 0. From the comparison between Tables 4 and 6 it can be
Deep percolation, actual transpiration and actual evaporation at time t = tend as a concluded that the percentage from the neglected evaporation is
percentage of the applied irrigation depth. mainly added to the actual transpiration. The percentage of
Discharge rate Soil (%) evaporation was 11.9% and the actual transpiration increased
1 1
Deep percolation 2lm h Loamy sand 15.4
Silt 6.1 Table 6
Silty clay loam 1.0 Deep percolation and actual transpiration at time t = tend as a percentage of the
1 1
4lm h Loamy sand 15.2 applied irrigation depth (soil evaporation is neglected).
Silt 5.9
Silty clay loam 0.9 Discharge rate Soil (%)
1 1 1 1
Actual transpiration 2lm h Loamy sand 72.7 Deep percolation 2lm h Loamy sand 16.4
Silt 80.8 Silt 6.3
Silty clay loam 86.8 Silty clay loam 1.1
4 l m1 h1 Loamy sand 72.9 4 l m1 h1 Loamy sand 16.0
Silt 80.9 Silt 5.9
Silty clay loam 86.7 Silty clay loam 1.1

Actual evaporation 2 l m1 h1 Loamy sand 11.9 Actual transpiration 2 l m1 h1 Loamy sand 83.6
Silt 13.1 Silt 93.7
Silty clay loam 14.2 Silty clay loam 101.1
4 l m1 h1 Loamy sand 11.9 4 l m1 h1 Loamy sand 84.0
Silt 13.2 Silt 94.1
Silty clay loam 14.2 Silty clay loam 101.1
S. Elmaloglou, E. Diamantopoulos / Agricultural Water Management 96 (2009) 1587–1595 1595

from 72.7 to 83.6%, while only a small increase in deep percolation Bhatnagar, P.R., Chauhan, H.S., 2008. Soil water movement under single surface
trickle source. Agric. Water Manage. 95 (7), 799–808.
from 15.4 to 16.4% is obtained. The physical explanation of this Ben-Asher, J., Phene, C.J., 1993. Analysis of surface and subsurface drip irrigation
difference is that when evaporation is neglected the soil water is using a numerical model. In: Subsurface Drip Irrigation—Theory, Practices and
more easily taken up by the plant roots. Application, California State University, Fresno, CA, pp. 185–202. CATI Pub. No.
92 1001.
Camp, C.R., 1998. Subsurface drip irrigation: a review. Trans. ASAE 41 (5), 1353–
4. Conclusion 1367.
Coelho, F.E., Or, D., 1996. A parametric model for two-dimensional water uptake
intensity by corn roots under drip irrigation. Soil Sci. Soc. Am. J. 60 (4), 1039–
A mathematical model which describes water flow under 1049.
subsurface drip lines, taking into account root water uptake, Cote, C.M., Bristow, K.L., Charlesworth, P.B., Cook, F.J., Thorburn, P.J., 2003. Analysis
evaporation of soil water from the soil surface and hysteresis in the of soil wetting and solute transport in subsurface trickle irrigation. Irrig. Sci. 22,
143–156.
soil water characteristic curve u(H) is presented. The model was
Elmaloglou, S., Diamantopoulos, E., 2008. The effect of hysteresis on three-dimen-
validated by comparison with an existing analytical solution and sional transient water flow during surface trickle irrigation. Irrig. Drain. 57, 57–
the Hydrus 2D model. The results showed that the numerical 70.
solution agrees very well with both the analytical solution and the Evett, S.R., Howell, T.A., Schneider, A.D., 1995. Energy and water balances for surface
and subsurface drip irrigated corn. In: Proceedings of the Fifth International
Hydrus 2D model. Microirrigation Congress, April 2–6, Orlando, FL, USA, pp. 135–140.
The results demonstrated that opportunities exist to improve Gardenas, A.I., Hopmans, J.W., Hanson, B.R., Simunek, J., 2005. Two-dimensional
the performance of SDI systems by adjusting system design and modeling of nitrate leaching for various fertigation scenarios under micro-
irrigation. Agric. Water Manage. 74, 219–242.
management strategies to account for differences in soil Feddes, R.A., Kowalik, P.J., Zaradny, H., 1978. Simulation of Field Water Use and Crop
hydraulic properties and discharge rate of line sources. The Yield. Simulation Monographs, Pudoc, Wageningen, The Netherlands.
numerical results showed that the soil wetting pattern mainly Hoogland, J.C., Feddes, R.A., Belmans, C., 1981. Root water uptake model depending
on a soil water pressure head and maximum extraction rate. Acta Horti. 119,
depends on soil hydraulic properties; that, for all simulation 123–136.
times, the average water content throughout the effective root Ismail, S.M., Zin El-Abendin, T.K., Wassif, M.A., El-Nesr, M.N., 2006. Wetting pattern
zone increased from the coarser to the finer soil; that at time simulation of surface and subsurface drip irrigation systems, II-Model valida-
tion and analyses. In: The 14th Annual Conference of Misr Society of Agricul-
t = tend deep percolation decreased from the coarser to the finer tural Engineering 24 (4). pp. 1035–1057.
soil; on the contrary, actual transpiration and actual soil Kool, J.B., Parker, J.C., 1987. Development and evaluation of closed-from expres-
evaporation increased from the coarser soil to the finer soil. sions for hysteretic soil hydraulic properties. Water Resour. Res. 23 (1), 105–
114.
In addition, the results showed that, for the same soil, at time
Lamm, F.R., Camp, C.R., 2007. Maintenance. In: Lamm, F.R., Ayars, J.E., Nakayama,
t = ti decreasing the discharge rate of line sources but main- F.S. (Eds.), Microirrigation for Crop Production. Design, Operation, And Manage-
taining the applied irrigation depth, the vertical and horizontal ment. Elsevier, Amsterdam, pp. 473–551.
components of the wetting front were increased; and at time Or, D., 1996. Drip irrigation in heterogeneous soils: steady state field experiments
for stochastic model evaluation. Soil Sci. Soc. Am. J. 60 (5), 1339–1349.
t = tend the discharge rate has no effect on the actual transpira- Philip, J.R., 1968. Steady infiltration from buried point sources and spherical
tion and actual evaporation and only a small effect on deep cavities. Water Resour. Res. 4 (5), 1039–1047.
percolation. Rubin, J., 1968. Theoretical analysis of two-dimensional, transient flow of water in
unsaturated and partly unsaturated soils. Soil Sci. Soc. Am. Proc. 32, 607–615.
The numerical results also showed that at time t = tend when the Schaap, M.G., Leij, L.J., 1998. Database-related accuracy and uncertainty of pedo-
daily potential transpiration is decreased the decrease of actual transfer functions. Soil Sci. 163, 765–779.
transpiration is almost equally distributed to the deep percolation Scott, P.S., Farquhar, G.J., Kouwen, N., 1983. Hysteretic effects on net infiltration.
Advances in infiltration, vol. 11–83. Am. Soc. Agric. Eng. Publication, St. Joseph,
and actual evaporation; and that at time t = tend when the soil MI, pp. 163–170.
evaporation is neglected the soil water is more easily taken up by Simunek, J., Sejna M., van Genuchten, M.T., 1999. The Hydrus-2D software package
the plant roots. for simulating two dimensional movement of water, heat and multiple solutes
in variably saturated media., Ver.2.0. Rep IGWMC-TPS-53, Int. Ground Water
Model. Cent., Colo. School of Mines, Golden.
Acknowledgments Skaggs, T.H., Trout, T.J., Simunek, J., Shouse, P.J., 2004. Comparison of Hydrus-2D
simulations of drip irrigation with experimental observations. J. Irrig. Drain.
Eng. ASCE 130 (4), 304–310.
Special and sincere thanks go to the two anonymous reviewers
Van Genuchten, M.T., 1980. A closed form equation for predicting the hydraulic
for their useful comments and suggestions. conductivity of unsaturated soils. Soil Sci. Soc. Am. J. 44, 892–898.
Warrick, A.W., Lomen, D.O., 1976. Time-dependent linearized infiltration, III: strip
References and disc sources. Soil Sci. Soc. Am. J. 40, 639–643.
Warrick, A.W., Lomen, D.O., Amoozegar-Fard, A., 1980. Linearized moisture flow
Ababou, R., 1981. Modélisation des transferts hydriques dans un sol en infiltration with root extraction for three dimensional, steady conditions. Soil Sci. Soc. Am. J.
localisée. Thèse de Docteur Ingénieur. Université de Grenoble. 44 (5), 911–914.

You might also like