Anda di halaman 1dari 13

Computers and Electronics in Agriculture 56 (2007) 161173

Predicting drift from field spraying by means of a 3D


computational fluid dynamics model
K. Baetens a, , D. Nuyttens b , P. Verboven a , M. De Schampheleire c ,
B. Nicola a , H. Ramon a
a

BIOSYST-MeBioS, Catholic University of Leuven, de Croylaan 42, 3001 Leuven, Belgium


Institute for Agricultural and Fisheries Research (ILVO), Scientic Institute of the Flemish Community, Unit Technology and
Food-Agricultural Engineering, Burg. Van Gansberghelaan 115, 9820 Merelbeke, Belgium
c Department of Crop Protection, University Ghent, Coupure Links 653, 9000 Ghent, Belgium
Received 14 April 2006; received in revised form 25 January 2007; accepted 26 January 2007

Abstract
In order to investigate and understand drift from field sprayers, a steady state computational fluid dynamics (CFD) model was developed. The model was developed in 3D in order to increase the understanding of the causes of drift: a deviation in the wind direction
cannot be captured by a 2D approach, the wake behind a wind screen is not symmetrical, the effects of a changed nozzle orientation
may not be symmetrical. The models accuracy was validated with field experiments carried out according to the international standard ISO 22866. A field sprayer with a spray boom width of 27 m and 54 nozzles (Hardi ISO F110-03 at 3 bar) was driving at 2.22 m/s
over a flat pasture. During the experiments the wind direction was perpendicular to the tractor track. The model explained the variation
in drift replicates during each single field experiment through varying boom height (0.30.7 m), wind velocity (1.32.5 m/s), wind
deviation (18 to +18 ) from the direction perpendicular to the tractor track and injection velocity of the droplets (1727 m/s). Boom
movements had the highest impact on the variations in drift values (deviations in drift deposits of 25%), followed by variation in wind
velocity (deviations in drift deposits of 3%) and injection velocity of the droplets (deviations in drift deposits of 2.5%). Wind deviation from the direction perpendicular to the tractor track had a reducing effect on the drift values (deviations in drift deposits of 2%).
Small variations in driving speed had little influence on drift values. Near drift (<5 m) is predicted well by the model but the increased
complexity compromised the predictions at greater distances. The model will be further developed in order to improve far drift prediction. Dynamic simulations will be performed and the model for turbulent dispersion will be optimized. The model did not require
calibration.
2007 Elsevier B.V. All rights reserved.
Keywords: Spray drift; Field sprayer; Computer simulation; Model; Sensitivity analysis; Crop protection

1. Introduction
Drift or the unintentional deposit of pesticides next to the application field during spraying may cause many problems,
including damage to neighboring crops, ecosystems, water ways and human health. Especially in Flanders which is
known to have an intensive agriculture and relatively small fields that are scattered and close to living area, the need

Corresponding author. Tel.: +32 16 320588; fax: +32 16 322955.


E-mail address: katrijn.baetens@biw.kuleuven.be (K. Baetens).

0168-1699/$ see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.compag.2007.01.009

162

K. Baetens et al. / Computers and Electronics in Agriculture 56 (2007) 161173

for a good understanding of drift from field sprayers is a high priority. A review about the current knowledge of drift
has been presented by Gil and Sinfort (2005).
There are a number of good field drift models available (Holterman et al., 1997; Mokeba et al., 1997; Zhu et
al., 1994; Miller and Hadfield, 1989; Thompson and Ley, 1983). However, none of the above models investigate the
variability found in drift values during field experiments. The variability of drift can be quite high during one experiment (Nuyttens et al., 2006b). This variability can be significant in order to make a realistic drift risk assessment.
Some of the earlier developed models are only 2D, whereas the arrangement of the nozzles and the air entrainment requires a 3D approach. Therefore these models only start modelling the droplet tracks at a rather undefined
position under the nozzle (Zhu et al., 1994; Miller and Hadfield, 1989). The 2D/3D model of Holterman et al.
(1997) accounts for 3D nozzle flow and air entrainment. However, the droplet flow model is not fully coupled to
the wind flow model, resulting in a pragmatic approach to the entrainment problem. A practical pesticide application tool for farmers was developed by Zhu et al. (1995), but this tool is based on a 2D model verified against
results obtained by droplets of a uniform diameter and validated in a wind tunnel, so not including the effect of variability in environmental parameters nor including the effect of a non-uniform droplet size distribution produced by
nozzles.
The presented work aims to contribute to the interpretation of the already developed models by developing a 3D fully
mechanistic model constructed in CFX 5.7 (Ansys CFX, Canonsburg, PA). A 3D approach was chosen because not all
flow problems have a symmetrical solution. A deviation in the wind direction cannot be captured by a 2D approach: the
wake behind a wind screen is not symmetrical and the effects of a changed nozzle orientation may not be symmetrical.
The Reynolds averaged NavierStokes equations with a k model for turbulence accounting for the flow field model
were fully coupled to a Lagrangian particle tracking model. A Lagrangian approach was chosen due to its particular
advantage of straightforwardly accounting for droplet trajectory crossing (Chen and Pereira, 1996). The drift model
was validated by means of field experiments. During field experiments it is impossible to keep wind velocity, wind
direction, boom height and tractor speed constant. Moreover, it is difficult to obtain realistic input data for injection
velocity of droplets. Therefore an investigation on the influence of the temporal variation of these parameters during
experiments was performed.

2. Materials and methods


2.1. Field experiments
The model was developed to simulate field experiments performed in the summer of 2005 in Merelbeke, Belgium
(ILVO, Nuyttens et al., 2006b). These experiments were all carried out according to the international standard ISO 22866
(2005) for field drift measurements. In these experiments a trailed field sprayer (Hardi Commander Twin Force) with a
boom width of 27 m was spraying a 100 m distance with a nominal speed of 2.22 m/s over a pasture. No air assistance was
used. The wind direction was approximately perpendicular to the driving direction of the tractor. The sprayed liquid consisted of water mixed with Brilliant Sulfo Flavine (BSF), a fluorescent tracer. The nozzle distance and boom height were
0.5 m. The nozzle type used was a Hardi ISO 110 03 standard flat fan nozzle at 300 kPa (1.2 l/min) resulting in an application volume of 180 l/ha. The nozzles were placed in such a way that the longest axis of the cross section of the spray
made an angle of 8 with the longitudinal axis of the boom sprayer. Samples of filter paper (0.25 m 0.25 m) were put at
0.5, 1, 2, 3, 5, 10, 15 and 20 m from the directly sprayed zone in order to investigate drift at various distances down wind
from the outer nozzle of the boom. This sample row was replicated three times during each run with a distance of 10 m
between each replicate. Drift (vol.%) in each sampling point was calculated by dividing the mass deposit at the sampling
points by the measured mass deposit in the directly sprayed zone. During each experiment the wind velocity and direction
at 1.5 and 3.25 m height were measured by means of ultrasonic anemometers together with air temperature and relative
humidity.
Three experiments (in three replicates) at different meteorological conditions were performed and used for model
validation. Experiment 1 had a reference (at 1.5 m height) wind velocity of 2.0 m/s, Experiment 2 3.1 m/s and
Experiment 3 3.9 m/s. Table 1 details the meteorological circumstances of each experiment. The drift deposit results
of the experiments are in the same range as found by Holterman et al. (1997).

K. Baetens et al. / Computers and Electronics in Agriculture 56 (2007) 161173

163

Table 1
Meteorological values during each experiment

Wind deviation ( )
Deviation range ( )
Relative humidity (%)
Temperature ( C)
Velocity at 1.5 m (m/s)
Velocity at 3.25 m (m/s)
Maximum velocity at 1.5 m (m/s)
Minimum velocity at 1.5 m (m/s)

Experiment 1

Experiment 2

Experiment 3

18.08
1.38
79.99
16.19
2.00
2.47
2.22
1.30

18.50
20.00
61.79
22.03
3.12
3.67
3.80
2.64

3.13
12.63
68.06
13.61
3.94
4.62
3.94
3.10

A negative deviation means the tractor is driving down wind.

2.2. Model formulation


The model uses a 3D CFD (computational fluid dynamics) approach to predict drift. The CFD code used in this work
is ANSYS CFX 5.7 (Ansys, Inc., Canonsburg, PA, USA). The domain width was set to 60 m (20 m of sampling length
and 27 m of boom width and some space for the wind profile to settle). The length was set to 100 m (representing the
driving distance covered by the tractor). The 3D model domain (see Fig. 1) consisted of a base plane (called pasture
plane) of 60 m 100 m representing the pasture. The height of the domain has to be sufficient to model the boundary
layer. This height was determined to be 6 m by means of numerical experiments and verification against the methods
and results of Blocken et al. (2007).
2.3. Droplet spray model
For discrete droplets travelling in a continuous fluid medium, the acting forces on the droplet that affect the droplet
acceleration are due to the difference in velocity between the droplet and the fluid and due to the displacement of the
fluid particle. Neglecting all the other forces except the drag and buoyancy forces, the following equation of motion
for the droplet was used:
dvdi
1
1
(1)
= d 2 cd |vi vdi |(vi vdi ) + d 3 (d )gi
dt
8
6
where md is the droplet mass (kg), d the droplet diameter (m), suffix i represents the direction (i = x, y or z), i
the continuous fluid velocity including a turbulent component that implements the turbulent dispersion (m/s), di the
discrete droplet velocity (m/s), cd the drag coefficient, d the discrete droplet density (kg/m3 ), the air density (kg/m3 ),
gi the gravitational acceleration (m/s2 ), and t is the time (s).
The drag coefficient (cd ) in Eq. (1) was calculated using the IshiiZuber drag model (Ishii and Zuber, 1979). It is
an empirical correlation based on the shape and state of the particles. At sufficiently small Reynold numbers droplets
behave as solid spherical particles, the drag coefficient for this regime will be referred by as cd,sphere . At larger Reynolds
md

Fig. 1. Scheme of the rectangular domain used as the drift modeling area. The black arrows represent the cross wind. The white arrow represents
the driving direction. The position of the nozzles (27 in total) is indicated by the white line on the bottom of the model.

164

K. Baetens et al. / Computers and Electronics in Agriculture 56 (2007) 161173

numbers droplets become first ellipsoidal in shape (drag coefficient will be referred by as cd,ellipse ) to finally become
spherical cap shaped (drag coefficient will be referred by as cd,cap ). The drag coefficient for spherical fluid droplets is
based on the Schiller Naumann drag coefficient (Schiller and Naumann, 1933) using a Reynolds mixture number for
incorporating the density of the droplets:
cd,sphere =

24
(1 + 0.15Re0.687
)
m
Rem

where Rem is a mixture Reynolds number based on a mixture viscosity, defined by


Rem =

|vi vdi |d
m

and m is the mixture viscosity (kg/(m s)) defined by


m
= (1 rd )2.5
c
c is the viscosity of the continuous phase (kg/(m s)), rd the discrete phase volume fraction and * is defined by
=

d + 0.4c
d + c

with d is the viscosity of the discrete phase (kg/(m s)).


The drag coefficient for densely distributed ellipse shaped particles is defined as
cd,ellipse = E(rd ) cd
E(rd ) defined by
E(rd ) =

1 + 17.67f (rd )6/7


18.67f (rd )

f (rd ) =

c
(1 rd )1/2
m

and

cd defined by
cd = 23 E0 1/2
where E0 or the Eotvos number is known as the dimensionless group through which the drag coefficient of ellipse
shaped droplets is determined. It is defined by
E0 =

g  dp2

with g is the gravitational acceleration (m/s2 ),  the density difference of the two phases (kg/m3 ) and is the surface
tension coefficient (N/m).
The drag coefficient for densely distributed cap shaped particles is defined as
cd,cap = (1 rd )2 cd
with cd = 8/3.
The CFX code pulls through an automatic regime selection by the following procedure:
cd = cd,sphere

if cd,sphere cd,ellipse

cd = min(cd,ellipse , cd,cap )

if cd,sphere < cd,ellipse

Most drift models use the model of Schiller and Naumann (1933) to calculate the drag coefficient. The benefit of
this model is that the effect of the density of the droplets and the shape of the droplets on the droplet track is integrated
in the calculation of the droplet track.

K. Baetens et al. / Computers and Electronics in Agriculture 56 (2007) 161173

165

The particle displacement (xi ) is calculated using forward Euler integration of the particle velocity over time step,
t:
xin = xio + vdi t
where the superscripts o and n refer to the old and new values, respectively. The particle velocity at the beginning
of the time step is assumed to prevail over the entire step. At the end of each time step, the new particle velocity is
calculated from Eq. (1). As the velocity fluctuations of the air flow are not available within the k model it is the task
of a turbulent dispersion model to regenerate accurate velocity fluctuation. The model for turbulent dispersion used in
this study was developed by Gosman and Ioannides (1981).
The source/sink term Su in Eq. (2) represents the integrated effects of momentum exchange with the continuous phase.
These Eulerian variables are calculated from the Lagrangian droplet variables by volume averaging the contributions
from all individual droplets within the cell volume. The coupling term is given as
Su =

nd
nd
d
dvdi
1 
1 
md
(md vdi ) =
Vcell
dt
Vcell
dt

(2)

where nd is the number of droplets in the cell and Vcell is the cell volume (m3 ). The first term on the right-hand side
of Eq. (2) results from the interaction forces between the two phases without any phase change. A more in depth
mathematical description can be found in Delele et al. (2005).
2.4. Atmospheric airow model
Since the air flow of interest takes part of a bigger flow field (the atmosphere), it is of high importance that proper
boundary conditions are set in order to simulate a continuum between modelled field area and the atmosphere.
The wind flow is modelled by introducing a logarithmic wind profile above the crop at the upwind side of the
pasture. The logarithmic wind profile is given by the following equation:


u
z + dr
u(z) =
ln

dr
where u(z) (m/s) is the wind velocity at a height z (m), u* the friction velocity (m/s). the von Karman constant (0.41)
and dr the roughness length (m). The roughness length of this profile is fitted to the experimental wind velocity profile
of the experiment (see Fig. 2 for an example). The deviation of the cross wind direction was measured and taken into
account in the model. Due to the small height of the pasture grass, the wind profile inside the crop canopy has not been
modelled. Relative humidity and temperature were assumed to be constant throughout the simulations. The tractor was
driving in a crosswind at 2.22 m/s. This is simulated by introducing a moving coordinate system in which the pasture
plane was set to be a moving wall with a velocity of 2.22 m/s in the opposite direction of the driving direction. At the
inlet planes an apparent wind of 2.22 m/s in the same direction was added.
The kinetic energy and dissipation parameters are implemented at the inlet boundaries according to the theory of
atmospheric boundary layers (Blocken et al., 2007):
k=

u2
,
C0.5

u3
(z + dr )

with k is the kinetic energy (m2 /s2 ). C is an internal model constant (0.09) and is the eddy dissipation velocity
(m3 /s2 ).
The airflow above the field was solved by means of the Reynolds averaged NavierStokes equations with the k
model of turbulence (Wilcox, 2000) and source terms. The pasture was assigned a roughness length of 0.01 m, the
value for grass lands (Tieleman, 2003).
2.5. Nozzle characteristics
The droplet diameter distribution of the ISO 110-03 nozzle (at 300 kPa) was obtained from experiments with a Phase
Doppler Particle Analyser (Aerometrics Inc.) in a conditioned room (Nuyttens et al., 2006a). The droplet diameter

166

K. Baetens et al. / Computers and Electronics in Agriculture 56 (2007) 161173

Fig. 2. Logarithmic wind profile at the inlet planes of the model (solid line). The dots represent the experimental values of the wind speed at 1.5 m
(2 m/s) and at 3.25 m (2.47 m/s). The obtained roughness length (dr ) was 0.056 m.

distribution was fitted by a Rosin Rammler model (Fig. 3) with a spreading factor of 3.2441 and a diameter de at
which the cumulative mass fraction becomes 1/e. The value of de was 310 m. The liquid sheet break-up length of
the nozzle was 0.023 m (Bayvel and Orzechowski, 1993). The randomized horizontal injection position of the droplets
was attained by distributing the droplets at random over the ellipse area of the spray cross section at a distance from
the nozzle equal to the break-up length. The vertical injection velocity of the droplets was assumed constant and equal
to 20 m/s for all droplets (Sidahmed et al., 2005). Sidahmed et al. (2005) measured a droplet velocity in the range of
1422 m/s with an average of 17 m/s at 0.04 m from the orifice of a flat fan nozzle ISO F 110-03 at 200 kPa. Droplet
velocity was found to depend on droplet diameter. For this work (working pressure of 300 kPa), the average velocity
was augmented with a factor of (3/2)1/2 based on Bernouillis law. A constant velocity injection value of 20 m/s was
therefore used, as was the case in the models of Zhu et al. (1995) and Holterman et al. (1997). The other injection
velocity components were determined from the fan angle (110 ), its shape and the droplets starting position in the
ellipse.
2.6. Numerical procedure
The model was solved numerically using the finite volume method. The simulations were performed on a P4 2.6 GHz
PC with 1 Gb RAM requiring an average of 65,000 CPU seconds for each simulation. Unstructured tetrahedral volume

Fig. 3. Rosin Rammler distribution fitted to experimental data of the droplet size distribution of an ISO F110-03 nozzle working under a pressure
of 300 kPa, a spreading factor () of 3.2441 and a reference diameter (de ) of 310 m.

K. Baetens et al. / Computers and Electronics in Agriculture 56 (2007) 161173

167

Fig. 4. Predicted drift values (vol.%) obtained at different distances from the field by modeling 1000 dpn (droplets per nozzle), 10,000 dpn and
20,000 dpn.

elements were used with prismatic elements on the ground surface. Eight lakh elements were required for mesh
independent results. The drift values from a reference simulation with 27 nozzles and injecting respectively 1000,
10,000 and 20,000 droplets per nozzle are presented in Fig. 4. The drift values for a model using 1000, 10,000 and
20,000 droplets per nozzle were almost identical, with increasing relative differences with distance. A trade-off between
computational effort and accuracy was achieved at 10,000 droplets per nozzle.
Drift values from a reference simulation using 14, 27 and 54 nozzles are compared in Fig. 5. From Fig. 5 it becomes
clear that droplets from the outer nozzle sprays still contribute to drift. Indeed, the drift profile changes when modelling
the complete boom (54 nozzles) and when modelling only half a boom (27 nozzles). The far distance (>5 m) drift profile
is most affected, though considering the logarithmic scale of the graph minimal. The changes in the nearby drift profile
are small. A trade-off in favour of the use of computational effort was made and simulations were performed with 27
nozzles.

Fig. 5. Predicted drift values (vol.%) obtained at different distances from the field by modelling 14 nozzles, 27 nozzles and 54 nozzles.

168

K. Baetens et al. / Computers and Electronics in Agriculture 56 (2007) 161173

2.7. Parameter sensitivity study


The influence of the variation of boom height, tractor speed, injection velocity of the droplets and wind speed on
drift values was studied. While the model is formulated in steady state, only the effect of the vertical boom height was
investigated without dynamic boom movement. Langenakens et al. (1999) measured a maximum deviation of 0.39 m
for a boom moving at 2.22 m/s. Simulations were therefore performed with three different boom heights, namely 0.3,
0.5 and 0.7 m. Driving speeds were changed from 1.94 m/s (or 7 km/h) to 2.22 m/s (or 8 km/h) to 2.5 m/s (or 9 km/h),
according to the round off error of the tractor speed indicator. During the performance of one experiment, the wind
velocity at a height of 1.5 m was changing. For example, in Experiment 1, the minimally and maximally measured
wind velocity values were 1.3 and 2.2 m/s with an average value of 2.0 m/s. These three values are used to study the
effect of a changing wind velocity during one experiment. The tractor was not driving perfectly crosswind, a deviation
from this direction was measured during the experiment. This deviation was implemented in the model. The resulting
variation in drift obtained by the model was compared to the variation observed in each experiment. As explained
earlier, Sidahmed et al. (2005) measured a droplet velocity in the range of 1422 m/s 0.04 m from the orifice of a flat
fan nozzle ISO F 110-03 at 200 kPa. Hence, after augmenting these outer values with a factor of (3/2)1/2 , simulations
with droplet velocities of 17, 20 and 27 m/s were studied.
Eventually, the obtained drift curves were evaluated by calculating the surface under the drift curve. Using this
method the total drift of a simulation over a distance from 0 to 20 m was determined. The same method was used to
determine short distance drift (05 m) and far distance drift (520 m).
3. Results
In Fig. 6 the predicted trajectories of individual droplets during field spraying are presented. The gray scale represents
the diameters of the droplets. The conical and dense sprays can be recognized underneath the nozzles. Due to the cross
wind, smaller droplets tend to drift across the field boundary. The smaller the droplet, the further they drift. It is also
important to notice that all nozzles contribute to the far distance drift (>5 m), while the near distance drift (<5 m) is
mainly due to larger drops from the nozzles closest to the field boundary.
In Fig. 7 the model predictions are compared to the measured drift profiles for each experiment. The difference
between experiments is mainly caused by variations in wind velocity. The predicted near distance drift (<5 m) corresponds well with the experimental values. At far distance, the model always predicts significantly lower values than
in the experiment. In average terms, the total measured nearby (<5 m) drift increased from 9.1% over 14.4% to 19.2%
for an increase in wind velocity from 2.0 to 3.1 to 3.9 m/s, respectively. The corresponding far distance (>5 m) drift
increased from 1.35% over 4.32% to 12.2%. The model predictions for these conditions were 8.03%, 11.8% and 16.1%

Fig. 6. Visualisation of the model output. The black line represents the theoretical boundary of the field. Model parameters: boom height of 0.5 m,
27 nozzles (ISO F110-03 at 3 bar), driving speed of 2.22 m/s (or 8 km/h).

K. Baetens et al. / Computers and Electronics in Agriculture 56 (2007) 161173

169

Fig. 7. Validation of Experiment 1 (a), Experiment 2 (b) and Experiment 3 (c). Bars are representing minimum and maximum values of the
replicates of each experiment. See Table 1 for the meteorological details during each experiment.

for the near distance and 0.5%, 1.54% and 2.8% for the far distance drift. There is an underestimation of drift by the
model for all the experiments, this underestimation is bigger at far distances. In the experiment, higher wind velocities
at 1.5 m imply more drift. This trend is reproduced by the model.
4. Discussion
The main discrepancy between the model and the experimental values exist in the far distance region, i.e. for drift
further than 5 m from the field. Other field drift models predict drift up till maximum 6 m (Holterman et al., 1997; Miller
and Hadfield, 1989; Mokeba et al., 1997). The model of Zhu et al. (1995) predicts drift up till 200 m. However, this
model was never validated against field experiments and describes the situation only for one single droplet. A first reason
for the discrepancy between experimental and model results is due to the models simplification for computational
reasons. The model uses 10,000 drops per nozzle and 27 out of a total of 54 nozzles on the boom. Figs. 4 and 5 show
that increasing the number of nozzles in the simulation contributes to the increase of far distance drift predictions. The
number of modelled nozzles has little effect on short distance drift.
Secondly, the simulation employed a simplified model of turbulent dispersion in an EulerianLagrangian approach
(Gosman and Ioannides, 1981), which has computational advances and has been widely used in the analysis of Klose
et al. (2001) comparing different isotropic dispersion approaches to turbulent dispersion, the choice of model did not
affect overall model performance for polydisperse systems. The drawback of all these approaches is that temporal

170

K. Baetens et al. / Computers and Electronics in Agriculture 56 (2007) 161173

Fig. 8. Sensitivity analysis of predicted drift values of Experiment 1. The input parameters are 2 m/s for wind velocity at 1.5 m with a variation
from 1.3 to 2.27 m/s for (a). The wind deviation in all graphs but (b) is 18 . In (b) there is a variation from 18 to 18 . The tractor is driving at
8 km/h with variation from 7 to 9 km/h in (c). The standard boom height is 50 cm with a variation from 30 to 70 cm for (d), and the injection velocity
of the droplets is 20 m/s with variation between 17 and 27 m/s in (e).

and directional correlations are not brought into account. To overcome this limitation, other formulations are required
including a different type of turbulence model (Reynolds stress models) for the continuous phase (Chen and Pereira,
1996). This would lead to a larger computational complexity and will be subject of future investigation.
The remaining discrepancy between model and experiment is believed to be mainly due to experimental variability
and uncertainty of the model input parameters. The experimental variability is large as shown in Fig. 7 by the error
bars. A sensitivity analysis was performed by means of model simulations to determine whether the observed model
discrepancy can be explained. The effect of the different factors is given in Fig. 8ae. The resulting total uncertainty
on the model prediction is given in Fig. 9 for the three experiments.
Wind velocity variations and wind direction are influenced by the surrounding topography (Raupach and Finnigan,
1997). Little is known about the quantitative influence of the topography on wind velocity profiles. The outer velocity
values at 1.5 m measured during Experiment 1 are tested for their effect on drift deposition using the model (Fig. 8a).
When the wind speed is reduced from 2 to 1.3 m/s the effect is quite considerable (in total 2.7% less drift with a lower
windspeed). An increase from 2 to 2.2 m/s has little or no effect on the drift values (an increase of 0.05% with higher
wind speed). The same range of variability is found by the model of Holterman et al. (1997). Experimental field results,
however, have a dynamic variability as well.
The influence of wind deviation is very little documented. In Experiment 1 the wind deviation was 18 upwind,
with less than 2 variation in wind direction during the experiment. Fig. 8b shows that the highest drift values occur
when the tractor is driving perfectly perpendicular to the wind direction. This is as expected because the cross wind

K. Baetens et al. / Computers and Electronics in Agriculture 56 (2007) 161173

171

Fig. 9. Comparison of modeled and experimental drift ranges at different distances. Dark bars represent the minimum and maximum values of the
experiments. Light bars represent modeled drift values with a boom height of 0.3 m (lower bars) and 0.5 m (highest bars). Figure (a) represents
Experiment 1, (b) Experiment 2 and (c) Experiment 3. See Table 1 for the meteorological details during each experiment.

drag forces become relatively less important when there is a deviation from the perpendicular wind direction. Fig. 8c
shows that small variations in driving speed have little or no influence on the variation in drift. This was also found by
Teske et al. (2001). It is reported in Holterman et al. (1997) that higher tractor speeds lead to higher drift depositions.
This trend cannot be seen in this study due to the small tractor velocity range that was investigated.
Fig. 8d shows that boom height has by far the most influence on drift values. A boom height of 0.3 m decreases
total drift by 8.4%, compared to a boom height of 0.5 m. A boom height of 0.7 m increases total drift by 17.02%.
These results are supported by the study of Herbst and Wolf (2001), who found for tractor mounted sprayers a range
of 1022% for coefficient of variation of spray deposit distribution.
Sidahmed et al. (1999) found by experimental investigation as well as theoretical considerations that smaller droplets
have a smaller injection velocity. Smaller droplets are more prone to drift. Fig. 8e shows that a lower injection velocity
results in a smaller amount of predicted short distance drift. Compared with an injection velocity of 20 m/s, there is
1% more short distance predicted drift with an injection velocity of 17 m/s. With an injection velocity of 27 m/s there
is 1.5% less short distance drift. Far distance drift remains more or less constant with different injection velocities
(1.44%, 1.20% and 1.16% short distance drift for injection velocities of 17, 20 and 27 m/s, respectively).
Fig. 9 shows that accumulated variation due to the above factors results in a model uncertainty comparable to that
of the experiment. Although, bringing into account this variation the model versus experimental discrepancy at far
distance is not fully captured. For this aspect, improved models for turbulence and near-nozzle characteristics are
necessary.

172

K. Baetens et al. / Computers and Electronics in Agriculture 56 (2007) 161173

Finally, inherent dynamic variation during experiments could only be explained fully by simulating field spraying
in a transient mode. Some experimental data on spatial and temporal variability in drift experiments have been made
available by Van De Zande et al. (2006). Future studies should incorporate such dynamic behavior.
5. Conclusions
A steady state 3D CFD model for drift from field spraying was able to predict short distance drift. Reliable predictions
were obtained at distances up to 5 m from the field boundary for the experimental conditions studied. The prediction
for far distance drift suffers from the limitations of the turbulence model that was used and the simplified descriptors of
the nozzle characteristics. Both items require considerable further model development before advances can be made.
The effect of small variations of some of the machine and environmental input variables on the model predictions
was significant. The resulting model uncertainty was due to measurement uncertainty as well as natural variability.
Boom height was according to this study the most sensitive parameter, followed by wind direction and wind velocity.
Tractor speed and injection velocity had little influence on the drift values. To fully account for the temporal fluctuations
of boom height, wind direction, wind velocity, injection velocity and tractor speed, dynamic simulations are required.
Acknowledgement
The authors wish to thank the IWT-Vlaanderen (Projects IWT-20424 and IWT-040708) for financial support.
References
Bayvel, L., Orzechowski, Z., 1993. Liquid Atomization. Taylor & Francis, Washington, USA.
Blocken, B., Stathopoulos, T., Carmeliet, J., 2007. CFD simulation of the atmospheric boundary layer: wall function problems. Atmos. Environ. 41,
238252.
Chen, X.-Q., Pereira, J.C.F., 1996. Computation of turbulent evaporating sprays with well-specified measurements: a sensitivity study on droplet
properties. Int. J. Heat Mass Transf. 39, 441454.
Delele, M.A., De Moor, A., Sonck, B., Ramon, H., Nicolai, B.M., Verboven, P., 2005. Modelling and validation of the air flow generated by a cross
flow air sprayer as affected by travel speed and fan speed. Biosyst. Eng. 92, 165174.
Gil, Y., Sinfort, C., 2005. Emission of pesticides to the air during sprayer application: a bibliographic review. Atmos. Environ. 39, 51835193.
Gosman, A.D., Ioannides, E., 1981. Aspects of computer simulation of liquid fuelled combustors. AIAA Paper No. 81-0323.
Herbst, A., Wolf, P., 2001. Spray deposit distribution from agricultural boom sprayers in dynamic conditions. ASAE meeting presentation. Paper
No. 011054.
Holterman, H.J., Van De Zande, J.C., Porskamp, H.A.J., Huijsmans, J.F.M., 1997. Modeling spray drift from boom sprayers. Comput. Electron.
Agric. 19, 122.
Ishii, M., Zuber, N., 1979. Drag coefficient and relative velocity in bubbly, droplet or particulate flows. AICHE J. 25, 843855.
ISO 22866, 2005. International Standard. Equipment for crop protection. Methods for field measurement of spray drift.
Klose, G., Rembold, B., Koch, R., Wittig, S., 2001. Comparison of state-of-the-art droplet turbulence interaction models for jet engine combustor
conditions. Int. J. Heat Fluid Flow 22, 343349.
Langenakens, J.J., Clijmans, L., Ramon, H., De Baerdemaeker, J., 1999. The effects of vertical sprayer boom movements on the uniformity of spray
distribution. J. Agric. Eng. Res. 74, 281291.
Miller, P.C.H., Hadfield, D.J., 1989. A simulation model of the spray drift from hydraulic nozzles. J. Agric. Eng. Res. 42, 135147.
Mokeba, M.L., Salt, D.W., Lee, B.E., Ford, M.G., 1997. Simulating the dynamics of spray droplets in the atmosphere using ballistic and random-walk
models combined. J. Wind Eng. Ind. Aerod. 67, 923933.
Nuyttens, D., De Schampheleire, M., Steurbaut, W., Baetens, K., Verboven, P., Nicola, B., Ramon, H., Sonck, B., 2006a. Experimental characterization of agricultural sprays using laser techniques. Aspects of applied biology 77, International advances in pesticide application,
179188.
Nuyttens, D., Brusselman, E., De Schampheleire, M., Steurbaut, W., Baetens, K., Verboven, P., Nicola, B., Ramon, H., Sonck, B., 2006b. Experimental study of factors influencing the risk of drift from field sprayers. Part 1. Meteorological conditions. Aspects of applied biology 77,
International advances in pesticide application, 321330.
Raupach, M.R., Finnigan, J.J., 1997. The influence of topography on meteorological variables and surfaceatmosphere interactions. J. Hydrol. 190,
182213.

Schiller, L., Naumann, A., 1933. Uber


die grundlegenden berechungen bei der schwerkraftaufbereitung. Vereines Deutscher Ingenieure 77, 318320.
Sidahmed, M.M., Brown, R.B., Darvishvand, M., 1999. Drop-size/velocity correlations at formation of sprays from fan nozzles. Trans. ASAE 42,
15571564.
Sidahmed, M.M., Taher, M.D., Brown, R.B., 2005. A virtual nozzle for simulation of spray generation and droplet transport. Biosyst. Eng. 92 (3),
295307.
Teske, M.E., Valcore, D.L., Hewitt, A.J., 2001. An analytical ground sprayer model. ASAE Paper 01-1051.

K. Baetens et al. / Computers and Electronics in Agriculture 56 (2007) 161173

173

Thompson, N., Ley, A.J., 1983. Estimating spray drift using a random-walk model of evaporating droplets. J. Agric. Eng. Res. 28, 419435.
Tieleman, H.W., 2003. Roughness estimation for wind-load simulation experiments. J. Wind Eng. Ind. Aerod. 91, 11631173.
Van De Zande, J.C., Holterman, H.J., Michielsen, J.M.P.G., Stallinga, H., 2006. Temporal and spatial variability of spray drift around a sprayed
field. Aspects of applied biology 77, International advances in pesticide application, 295302.
Wilcox, D.C., 2000. Turbulence Modelling for CFD. DCW Industries, p. 314.
Zhu, H., Reichard, D.L., Fox, R.D., Brazee, R.D., Ozkan, H.Z., 1994. Simulation of drift of discrete sizes of water droplets from field sprayers.
Trans. ASAE 37, 14011407.
Zhu, H., Reichard, D.L., Fox, R.D., Ozkan, H.E., Brazee, R.D., 1995. DRIFTSIM, a program to estimate drift distances of spray droplets. Appl.
Eng. Agric. 11, 365369.

Anda mungkin juga menyukai