J Biosystemseng 2016 10 010

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4

Available online at www.sciencedirect.com

ScienceDirect

journal homepage: www.elsevier.com/locate/issn/15375110

Special Issue: Spray Drift Reduction


Research Paper

Development and validation of a 3D CFD model


of drift and its application to air-assisted
orchard sprayers

Ashenafi T. Duga a,*, Mulugeta A. Delele a, Kris Ruysen c,


Donald Dekeyser b, David Nuyttens b, Dany Bylemans a,c,
Bart M. Nicolai a, Pieter Verboven a,**
a
KU Leuven e University of Leuven, Department of Biosystems, MeBioS, De Croylaan 42, 3001 Leuven, Belgium
b
Institute for Agricultural and Fisheries Research (ILVO), Technology and Food Science Unit, Agricultural
Engineering, Burg. Van Gansberghelaan 115, 9820 Merelbeke, Belgium
c
Research Station for Fruit Growing (PCFruit), Department of Ecology, Fruittuinweg 1, 3800 Sint-Truiden, Belgium

article info
Pesticides play an important role in providing high crop yields by minimising the risks
Article history: associated with pests but some of the sprayed product may move beyond the intended
Published online xxx target and result in drift. Modelling approaches can help understand the behaviour of
spray drift using computer simulations. However, modelling drift from orchard spraying
presents particular challenges: (1) the moving spray interacts with the canopy before
Keywords: drifting outside the target area; (2) the vertical wind profile in the orchard is different to
Drift reduction neighbouring fields where there is different vegetation; (3) the moving air jet from the air-
Nozzle assistance cannot be ignored because the airspeed of the fan is usually higher than the
Tree model wind speed.
Lagrangian model This work presents a three-dimensional (3D) computational fluid dynamics (CFD) model
Fan speed of spray drift from orchard sprayers that considers tree architecture, canopy wind flow and
Plant protection product the movement of the sprayer to calculate sedimenting and airborne drift; thus tackling the
challenges listed above. The model was validated against drift measurements from an
apple orchard with different nozzles arrangements. The model was then used to evaluate
the effect of drift reducing nozzles and fan airspeed on drift. The model predicted that drift
reducing nozzles reduced the drifting distance by 50%, but increased near-tree ground
deposition. This increase in ground deposition can be avoided whilst retaining the
reduction in the drifting distance, by using a combination of drift reducing and standard
nozzles. A reduced sprayer airflow can further reduce drift.
© 2016 IAgrE. Published by Elsevier Ltd. All rights reserved.

* Corresponding author.
** Corresponding author.
E-mail addresses: ashenafitilahun.duga@biw.kuleuven.be (A.T. Duga), pieter.verboven@biw.kuleuven.be (P. Verboven).
http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
1537-5110/© 2016 IAgrE. Published by Elsevier Ltd. All rights reserved.

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
2 b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4

drifting plume, are highly influenced by meteorological con-


Nomenclature ditions and data collection is labour and time consuming
(Gregorio et al., 2014). Light detection and ranging (LIDAR)
AAS Atomic Absorption Spectrometer
techniques have been used to some extent to monitor the
CFD Computational Fluid Dynamics
airborne spray drift from aerial and ground sprayers (Gregorio,
ks Equivalent sand grain roughness height
Rocadenbosch, Sanz, & Rosell-Polo, 2015; Miller, Saliyani, &
LIDAR Light detection and ranging
Hiscox, 2003; Stoughton, Miller, Yang, & Ducharme, 1997).
NRMSE Normalised-Root-Mean-Squared-Error
However, this technique is mostly used to study the move-
PDPA Phase Doppler Particle Analyser
ment and dispersion of the pesticide plumes qualitatively and
RMS Root-Mean-Square
it is only recently that some researchers have attempted to
3D Three dimensional
quantify droplet concentration in spray clouds using LIDAR
URANS Unsteady Reynolds averaged NaviereStokes
(Gregorio et al., 2015, 2014; Khot et al., 2011).
USDA U.S. Department of Agriculture
The high temporal and geographical variability of most of
USEPA U.S. Environmental Protection Agency
the factors that affect pesticide drift (crop characteristics,
Xe,i Measured drift value at position i
equipment design and setup, field size and slope, and num-
Xm,i Model prediction drift value at position i
ber of treatments) even in the same country make it difficult
y0 Roughness length
to monitor drift using experiments alone. Properly validated
yþ Minimum dimensionless distance from the
modelling approaches allow a controlled parameter analysis
wall
of the spraying process, thus better understanding of the
contribution of each parameter. A few models have been
developed and validated in the past few decades to study
drift from aerial and ground spray applications (Baetens
1. Introduction et al., 2009, 2007; Egan, Bohnenblust, Goslee, Mortensen, &
Tooker, 2014; Kruckeberg, Hanna, Steward, & Darr, 2012;
Agrochemical sprays play a pivotal role in enhancing the Nsibande, Dabrowski, van der Walt, Venter, & Forbes, 2015;
productivity and quality of crops by minimising losses. How- Teske, Thistle, & Ice, 2003; Teske et al., 2002). Some of
ever, their benefits are not without risks. A significant portion these models are empirical models developed through curve
of the sprayed material moves beyond the intended target and fitting (Lazzaro, Otto, & Zanin, 2008; Rautmann, Streloke, &
this poses environmental, economic and health risks. These Winkler, 2001). The applicability of these models is there-
associated risks have been the subject of discussion among fore limited to the site and the conditions under which the
the scientific community for over half a century. The first data used to develop them were collected. They were also
comprehensive report on the physical principles and mea- mostly developed for a single pollution target. Holterman
surements of drift was published by Akesson and Yates (1964). and Van de Zande (2008, 2010) developed a cascade drift
They reported on drift damages to non-target susceptible model to predict the spatial and temporal distribution of
crops from aerial applications. Since then, the focus of drift pesticide drift into a network of interconnected water bodies.
research has broadened and it now includes bystander and The model computes spray drift to multiple water bodies
resident exposure (Butler Ellis, Lane, O'Sullivan, Miller, & using a generic drift function developed by multiple linear
Glass, 2010), surface water contamination and the protection regression of a large set of random scenarios obtained from
of aquatic life (FOCUS, 2007; Lee et al., 2013). the IDEFICS (IMAG program for Drift Evaluation for Field
According to the ISO (International Organization for Stan- sprayers by Computer Simulation) drift model (Holterman,
dardization) standard (ISO 22866, 2005), spray drift is defined Van de Zande, Porskamp, & Huijsmans, 1997). The other
as the quantity of plant protection product that is carried out group of models are called mechanistic models that are
of the treated area by the action of air currents during the developed based on a set of physical equations describing a
application process. The amount of pesticide drifted beyond process (Baetens et al., 2009, 2007; Teske et al., 2002, 2003). As
the treated area depends on spray application technique such, they are independent of site and the conditions under
(Duga, Dekeyser, et al., 2015; Van De Zande et al., 2008), which the data used to develop them were collected. How-
physiochemical properties of the sprayed material (Dorr et al., ever, the applicability of mechanistic models is limited by
2013; Hilz & Vermeer, 2013), canopy architecture (Duga, their complexity and high computational demand. The most
Dekeyser, et al., 2015) and meteorological conditions advanced and widely used among these is the AGDISP®
(Arvidsson, Bergstro € m, & Kreuger, 2011; Duga et al., 2015a, model developed by the U.S. Department of Agriculture
2015b). Field tests (Nuyttens, De Schampheleire, Baetens, & (USDA) Forest Service for aerial application. An additional
Sonck, 2007b) and wind tunnel experiments (Nuyttens, model, AgDRIFT was separately created and spawned from
Taylor, De Schampheleire, Verboven, & Dekeyser, 2009) are AGDISP® under a Cooperative Research and Development
used to understand the mechanisms that govern pesticide Agreement between the Spray Drift Task Force and its part-
drift, characterise the effect of the different factors involved ners, the US EPA and the USDA Forest Service. The AgDRIFT
and assess drift potential (De Schampheleire, Baetens, model itself consists of three application modules that may
Nuyttens, & Spanoghe, 2008; Donkersley & Nuyttens, 2011; be used to estimate downwind deposition of spray drift from
Salyani, Miller, Farooq, & Sweeb, 2013). However, they both aerial, ground boom and orchard/vineyard air blast (air-
use sampling techniques and the data obtained are not time assisted) applications. However, the orchard/vineyard air
resolved, they also consider a limited number of points in the blast module is based on empirical curve fits of data from a

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4 3

few orchard field trials which limits its applicability outside 2.1. Field experiment
the experimental conditions of the data from which it was
developed. Hence, there is a need to develop models that 2.1.1. Sprayer design
could be applied irrespective of the particular spraying situ- A cross-flow sprayer with Power-Take-Off (PTO) driven axial
ation. A new integrated CFD model of orchard sprayers that fans (DuoProp, BAB Bamps, Sint-Truiden, Belgium) (Fig. 1) was
incorporates the real tree architecture and a porous medium used for both the field tests and model development. One-sided
to represent the leaves and small branches was developed spraying (right side of the sprayer) was considered in the
and validated using field experiments (Duga et al., 2015a; experiment. The velocity distribution of the sprayer air-jet and
Endalew et al., 2010b). This model takes into account the the spray characteristics of the nozzles used were measured
actual tree architecture, the canopy wind profile and sprayer before the drift field tests and used as inputs for the CFD model.
air flow and computes the droplet trajectory from a moving A hot wire anemometer (air velocity transducer, model 8465,
sprayer using the Lagrangian particle tracking model. While TSI, Shoreview, MN, USA) was placed as close as possible to the
it was validated using on-target and ground deposition sprayer outlet to measure the airflow. Additional measure-
within a single row of an orchard (Duga et al., 2015a), it ments were performed using 3D ultrasonic sensors (model
currently does not predict drift from orchard sprayers. The 81000, Young, Traverse City, MI, USA) placed at 0.15 m
objective of this study was thus to develop and validate a CFD perpendicular to the outlet. The air flow measurements were
model to predict the sedimenting and air borne drift from taken at a horizontal interval of 0.05 m following the contour of
orchard sprayers. The model calculation of sedimenting drift the air outlet. Figure 1 shows the vertical profile of the
from an air assisted orchard sprayer in a typical apple or- measured 3D velocity components of the sprayer air-jet. Details
chard in Belgium was compared to field trials. This model of the air flow measurements are given in Dekeyser et al. (2013).
was then used to predict the effects of using drift reducing The total air flow rate estimated using the measured air
nozzles and altering fan speed on spray drift. velocities and the corresponding outlet area was 50,000
(m3 h1) and 40,000 (m3 h1) for the high and low fan speed
settings respectively. A total of eight standard Albuz ATR or-
2. Materials and methods ange and TVI 8002 drift reducing nozzles (Saint-Gobain Sol-

cera, Evreux, France) were fitted to one side of the sprayer
In the following sections the model and field trials used for using the arrangements described in the next section. These
validation are explained. To better understand the specific nozzles were operating at a pressure of 600 kPa producing a
application, this section starts with a description of the field spray with a volume median diameter of 155.8 mm and 380 mm,
experiment, including the sprayer and nozzles used and a respectively. The sprayer was operated at an application rate
description of the drift measurements in an apple orchard. For of 500 l ha1 and a driving speed of 1.67 m s1 spraying only
this application, the corresponding CFD model is then one side of the row.
explained. If another sprayer type, nozzle or orchard system is
considered, then the CFD model can be easily adapted to these 2.1.2. Nozzle arrangements
as explained by Duga, Dekeyser, et al. (2015), Duga et al. Drift from three different nozzle arrangements was analysed
(2015a, 2015b). using the cross-flow sprayer. The first arrangement was using

Fig. 1 e The three components of the outlet velocity of the cross-flow sprayer (Duoprop, BAB-Bamps, Sint-Truiden, Belgium)
used in the analysis operating at high (a) and low fan speeds (b): U ( , horizontal and perpendicular to driving direction), V
( , vertical upward) and W ( , horizontal in the driving direction). The outlet velocities were measured at 0.15 m from the
outlet area at different heights using 3D ultrasonic sensors (Dekeyser et al., 2013).

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
4 b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4

the standard Albuz ATR orange hollow cone nozzles at all eight to measure the sedimenting drift. However, single side
positions of the prayer which is represented here after by ATR, spraying was used on the inside of the last row, spraying
the second arrangement was using Albuz TVI 8002 yellow drift outward, for the purpose of model validation. This spraying
reducing nozzles at all eight nozzle positions which is repre- is considered to contribute the majority of drift downwind
sented here after by TVI and the last arrangement was a from the orchard but it should be noted that the measure-
combination of the two nozzles types (Albuz TVI 8002 yellow ments could not be considered a full drift test. Spraying was
nozzles used at the top three positions and standard Albuz ATR carried out using metal tracers cobalt, manganese and
orange hollow cone nozzles used at the bottom five positions) magnesium at an intended concentration of 4000 ppm.
which is represented here after by (ATR þ TVI). After spraying, the samplers from each sampling position
The spray characteristics of the nozzles were measured were collected and stored at 4  C in dark conditions. The
using a one-dimensional phase Doppler particle analyser (PDPA, samplers were then washed with 0.16 N HNO3 solution to
Aerometrics, USA) (Nuyttens, Baetens, De Schampheleire, & extract the concentration of the tracer. The diluted solution in
Sonck, 2007a). The measured particle size distributions of the the test tubes was shaken for a minute and then the samplers
different nozzles types were then fitted to a Rosin-Rammler were removed from the solution. The amount of metal tracer
distribution. However, the best fit to the Rosin-Rammler distri- collected on each sampler was analysed using a Varian
bution that was obtained for the Albuz TVI 8002 yellow drift SpectrAA 300 atomic absorption spectrometer (AAS) (Varian
reducing nozzles was not as good as for the standard Albuz ATR Inc., CA, USA). The spray depositions at different distance and
orange nozzles (Fig. 2). Hence, the measured (as opposed to height were calculated using the surface area of the samplers.
fitted) size distributions were used in the model for both nozzles The sedimenting drift at a given distance was then calculated
types. The Albuz TVI 8002 drift reducing nozzles used in this as a percentage of the total amount sprayed.
analysis have a 50% drift reduction according to the Belgian A scientific weather station (Campbell Scientific, UT, USA)
buffer zone regulation (Anon, 2004) (see Fig. 3). was placed at 50 m from the sampling position to monitor
wind velocity and temperature at three heights (1, 2 and 3 m),
2.1.3. Drift measurement relative humidity at 2 m, and wind velocity and direction at
The field experiments were conducted in an experimental 3.5 m. Wind speed and direction were measured at 10 m
orchard in October 2013 (pcfruit, Sint-Truiden, Belgium) height using a 3D ultrasonic anemometer (Metek GmbH,
containing three-year-old apple trees under the classical Elmshorn, Germany), at 10 Hz. The measured meteorological
training system. Experiments were performed with the three data (wind velocity and direction, temperature, relative hu-
nozzle arrangements discussed in the previous section using midity) were used as inputs to the model.
the cross-flow sprayer. The trees in the orchard were ar-
ranged North-South with an interplant spacing of 1 m. The
inter-row spacing was 3.2 m. Replicate measurements were 2.2. CFD model
done by spraying 3 rows of trees. Each spray configuration is
replicated three times resulting in three separate measure- A CFD orchard drift model was developed based on an existing
ments of drift under comparable weather conditions. The CFD model for predicting the on-target spray distribution in
trees used for the field trial were 2.6 ± 0.3 m high and orchards (Duga et al., 2015a). The model considers the archi-
1.4 ± 0.3 m wide. A measurement protocol that was prepared tecture of the trees, the canopy wind flow (including both the
in accordance to the ISO standard (ISO 22866, 2005) was used within-canopy and above-canopy wind flows up to 3 times the
canopy height) and the moving sprayer outlet with dedicated
spray nozzles. It then computes the tracks of representative
droplets of the nozzle size distribution from the nozzle to the
target, to non-target surfaces directly around the tree and the
ones remaining in the air. This model was validated with on-
tree measurements of deposition (Duga et al., 2015a). The
model considered trees within the bulk of the orchard and was
restricted to a small domain around a single tree and two
neighbouring trees. For drift, however, a larger domain needs
to be considered to predict the ground and airborne drift at
larger distances behind a side row of trees. A computational
domain of 40 m length and 50 m width was used in this work
to represent the area next to the row of three orchard trees
(Fig. 4). The atmosphere was considered to 12 m high to
include the lower part of the orchard boundary layer and the
maximum sampling position used during the field trials. The
3D architecture of the trees was developed from the coordi-
Fig. 2 e The droplet size distributions and the nate data collected during the field tests and used in the model
corresponding Rossin-Rammler distribution fits of the two simulations. Details on the development of the tree architec-
nozzle types. The broken lines represent the measured ture can be found in Endalew et al. (2011). Only the outlet of
size distributions and the solid lines represent the Rossin- the sprayer was represented in the model; using a rectangular
Rammler distribution fits. cross-section. The measured outlet velocity profile of the

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4 5

Fig. 3 e The drift sampling positions and trees used for the field experiments.

Fig. 4 e The dimensions of the computational domain and the boundary conditions used in the CFD model. Wind is in the
same direction as the direction of spraying.

sprayer was applied to this cross-section to represent the right CFD code of ANSYS-CFX (ANSYS, Inc., Canonsburg, PA, USA).
side of half of the sprayer. The airflow model computed the transient airflow pattern
The wind and the air flow from the sprayer were modelled from the sprayer and its interaction with the wind and trees as
using the unsteady Reynolds averaged NaviereStokes the sprayer drives along the row. The effect of wind was in-
(URANS) equations and the k2 turbulence model which tegrated into the model using a canopy wind profile which
were solved using the unstructured finite volume method in a was obtained from a series of steady Reynolds averaged

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
6 b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4

NaviereStokes (RANS) simulations over the computational Processing Unit) time of 83 h using three computing nodes on
domain to match the average measured wind velocity and a KU Leuven HPC Linux cluster each having 64 GB of Random
direction of each trial obtained by the 3D anemometer at 10 m Access Memory (RAM). The model was solved for the cross-
height above the canopy, according to the procedure flow sprayer in an apple orchard and validated using dedi-
explained by Endalew et al. (2009). A series of cyclic simula- cated field trials. It was then used to compare spray drift from
tions were carried out using the measured wind speeds to the three nozzle arrangements and two fan speeds. Simula-
determine a realistic canopy wind profile that was used in the tions were carried out with the cross-flow sprayer using the
model simulations. Each following simulation used the outlet droplet size distributions measured from the three nozzle
profiles of the previous simulation as inlet boundary condi- setups. These simulations were done for the same wind con-
tion. The procedure was repeated until the difference between dition of magnitude 3.0 (m s1) measured at 10 m height
the inlet and outlet profiles of consecutive simulations blowing in the direction of spraying. The model can easily be
became insignificant. A normalised root-mean-square (RMS) adapted to other sprayer types and training systems (Duga
residual of less than 106 was used as the convergence crite- et al., 2015a).
rion for the steady state simulations. The resulting canopy The prediction error of the model was analysed using the
profiles were then imposed as input profiles at the boundaries Root-Mean-Squared-Error q (RMSE). The RMSE was estimated
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P
of the domain depending on the wind direction. The URANS using the relation RMSE ¼ n1 ni¼1 ðXe;i  Xm;i Þ2 where Xe,i is the
model was then solved by superimposing the outlet velocity measured value and Xm,i is the model prediction at position i.
profile of the sprayer which was defined as a moving boundary The normalised RMSE values which were calculated using the
conditions at the driving speed of the orchard sprayer to the range of the measurements (maximum minus minimum
canopy wind profile. This transient airflow model used the values) were then used to assess the prediction accuracy of
steady canopy wind profile from the cyclic simulations as an the model.
initial condition. The turbulent boundary condition of the
sprayer airflow was defined using a turbulent intensity of 30%
and a length scale of 0.008 m (Delele et al., 2005). No-slip rough 3. Results
wall boundary conditions were used for the surfaces of the
tree branches (equivalent sand grain roughness height 3.1. Validation of the simulated drift curves
(ks) ¼ 0.006 m) and the bottom boundary of the domain
(roughness length (y0) ¼ 0.005 m) (Endalew et al., 2009). The The drift curves were validated for three different nozzle ar-
other computational boundaries were set as atmospheric rangements (ATR, TVI, ATR þ TVI) fitted to the cross-flow
pressure openings to allow movement of air into and out of sprayer. A 3.0 to 3.9 (m s1) magnitude wind was blowing
the domain. The resistance and turbulence effects of the from north-east when the drift measurements were done
leaves were modelled using closure models applied in the using ATR nozzles (Fig. 5a). The magnitude of the wind
porous domain around the branches (Wilson & Shaw, 1977). registered during the experiment with the TVI nozzles was
A Lagrangian particle tracking multiphase flow model was relatively more variable and ranged from 1.0 to 4.6 (m s1)
used to calculate the instantaneous position of the spray (Fig. 5b). This wind was originally blowing from north-east
droplets in the turbulent airflow field around the trees and in and later shifted to an easterly wind. The wind registered
the drift zone behind the trees (Delele et al., 2007). The model during the experiment with the ATR þ TVI nozzles had a
uses the measured nozzle and spray parameters (spray angle, magnitude ranging from 1.6 to 3.9 (m s1). This wind originally
liquid flow rate and pressure, nozzle size, droplet size distri- blew from north-east but later it changed direction to an
bution) to track the droplets. The accuracy of a Lagrangian easterly wind (Fig. 5c).
particle tracking model highly depends on the number of Figure 6aec shows a comparison of the experimental
particles injected (Graham & Moyeed, 2002). In this work, 3000 measurements and model predictions for these three nozzle
particles were injected per time-step based on the sensitivity arrangements. The black and red lines in these plots represent
study of Delele et al. (2007). The deposition of droplets on the the measured values and the model predictions, respectively.
leaves was modelled using a stochastic deposition model As can be seen from the sedimenting drift qualitative plots in
which is a function of the optical porosity of the trees Fig. 6, the model results were in a good agreement with the
(Endalew et al., 2010a). This model calculates the number of experimentally determined values. The model in general
droplets captured by the porous domain using the vertical predicted the trend of the drift curves well with some differ-
profile of the optical porosity. The vertical profile of optical ences in the degree of agreement among the three nozzle ar-
porosity was calculated from the leaf area density and width rangements. The model had a prediction error of 26%, 23% and
of the tree (Raupach, Woods, Dorr, Leys, & Cleugh, 2001). The 32% for the three nozzle arrangements: ATR, TVI and
computational domain was discretised using an unstructured ATR þ TVI, respectively. This prediction error is attributed to
tetrahedral mesh combined with prismatic layers near the the temporal variations in wind conditions, the effect of
ground. The initial mesh size and the smallest mesh size near branch movement and droplet evaporation which are not
the surface of the tree branches were selected based on the considered in the model. As can be seen from the wind rose
required minimum dimensionless distance from the wall (yþ) plot in Fig. 5, there was a difference in the dynamics of the
for the turbulent wall functions to be valid (Kuzmin, Mierka, & wind even among the field trials for the three nozzle ar-
Turek, 2007). This resulted in a total of 11,839,661 elements rangements. The wind registered when the field trials were
and 2,184,962 nodes to simulate the cross-flow sprayer for the conducted using the TVI drift reducing nozzles and the
different settings. The calculations took a total CPU (Central ATR þ TVI nozzles (Fig. 5b and c) was relatively more variable

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4 7

Fig. 5 e Wind rose plot showing the magnitude and direction of wind registered at 10 m height when a field trial was
conducted on an apple classical training system using the cross-flow sprayer fitted with three nozzle arrangements: (a) ATR,
(b) TVI and (c) ATR þ TVI.

both in magnitude and direction than the wind registered row of trees. Figure 7a shows the droplet track plot obtained
during the trial with the ATR nozzles (Fig. 5a). However, the when ATR nozzles were used on the cross-flow sprayer. As can
prediction error of the model for the field test with ATR noz- be seen from this plot, most of the droplets are still airborne
zles was higher than that with TVI nozzles. This could be 40 m behind the tree row. This can be explained by the droplet
mainly because of the difference in the droplet size distribu- size distribution generated by the Albuz ATR orange nozzles.
tion generated by these two nozzle types as well as the factors These nozzles are characterised by a high percentage of fine
given above. The Albuz ATR orange nozzles generated a droplets as shown in Fig. 2. The maximum droplet size obtained
higher percentage of fine droplets than the Albuz TVI yellow from standard Albuz ATR orange nozzles is 430 mm. These fine
nozzles. These fine droplets tended to stay airborne and were droplets remain suspended in air for a longer time and cause
more susceptible to dynamic wind effects than the mainly drift at greater distances.
coarse droplets produced by the Albuz TVI nozzles which The droplet track plots obtained when TVI nozzles were
reached the ground after a shorter time in air. This shows that used at all eight positions on the cross-flow sprayer is shown in
droplet size distribution plays a significant part in the complex Fig. 7b. It can be seen from this plot that a significant portion of
interplay between spray, canopy and meteorological param- the droplets sedimented to the ground within the first 20 m.
eters to decide the final fate of the spray and the extent of drift There are relatively few droplets airborne compared to when
from orchard sprayers. ATR nozzles are used. The mainly coarse droplets from TVI
nozzles deposited to the ground much quicker than the drop-
3.2. Effect of nozzle arrangement on drift lets from the standard Albuz ATR orange nozzles. 45% of the
spray droplets generated by the Albuz TVI 8002 drift reducing
Figure 7 presents droplet track plots, coloured according to nozzles have a diameter greater than the maximum droplet
droplet size, for the three nozzle arrangements. This plot shows size generated by the standard Albuz ATR orange nozzles
the transient position of the droplets up to 40 m behind the last (Fig. 2). Earlier, Duga et al., 2014 had carried out laboratory

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
8 b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4

Fig. 6 e Validation of the drift curves predicted by the CFD model for three different nozzle arrangements fitted to a cross-
flow sprayer: (a) ATR, (b) TVI (c) ATR þ TVI. The red lines represent the model predictions and the black lines represent the
measurements (Error bars denote standard deviation). The magnitude and direction of the wind measured at 10 m height is
shown in Fig. 5. Application rate was 500 l ha¡1 and driving speed was 1.67 m s¡1.

experiments and CFD simulations using these three nozzle ATR þ TVI combination, when compared to the TVI nozzles,
arrangements on the cross-flow sprayer to study the spray gave relatively low ground deposition close to the tree but drift
distribution around the tree. When the Albuz TVI 8002 drift was detected at a greater distance (Fig. 8c). It can also be seen
reducing nozzles were used, a relatively higher proportion of from this figure that there is an apparent deviation in the
droplets deposited on the ground before reaching the canopy spray droplet trajectory around the trees. This is due to the
than when the standard Albuz ATR orange nozzles were used. small number of trees considered (only three trees in a row) in
This could compromise the on-target deposition depending on the model. This also contributed to the differences between
the wind and canopy density. However, it has been shown for the measurements and model predictions in the first few
the configurations considered in this study that the two nozzle metres behind the trees as shown in Fig. 6. This problem could
types could be used together to limit their individual disad- be avoided in future simulations by considering more trees on
vantages and provide an improved distribution. either side of the row.
Figure 7c presents the droplet track plot obtained when Figure 9 quantitatively summarises the sedimenting drift
ATR þ TVI nozzle are used. This plot shows that the droplet up to a distance of 40 m behind the trees for the three nozzle
trajectory obtained for this configuration lies between the two arrangements. The TVI nozzles gave the highest percentage
nozzle types with relatively lower proportion of droplets still drift in the first 3 m behind the trees. From 3 m to 12 m, the TVI
airborne than the standard Albuz ATR orange nozzles. and ATR þ TVI nozzle configurations gave similar percentage
Figure 8 shows contour plots of the time integrated spray drift which is larger than that from the standard Albuz ATR
deposition on the ground at different horizontal distances orange nozzles. From 12 m to 40 m, the standard ATR nozzles
behind the tree row for the three nozzle arrangements ob- gave the highest percentage drift. The sedimenting drift from
tained from the CFD simulation. As can be seen from this the TVI nozzles and the ATR þ TVI combination dropped to
figure, ATR nozzles produced ground deposition up to 40 m <1% in the first 20 m. The ATR nozzles maintained a per-
behind the trees (Fig. 8a) but this reduced to only 20 m when centage drift of close to 2% up to 40 m behind the trees.
TVI drift reducing nozzles were used (Fig. 8b). However, the Attention should be paid to the difference in the percentage
TVI nozzles produced higher deposition closer to the tree. The drift among the different nozzle arrangements rather than the

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4 9

Fig. 7 e Droplet track plots showing the droplet trajectories up to 40 m behind the last tree row for an apple classical training
system sprayed with the cross-flow sprayer using three nozzle arrangements: (a) ATR, (b) TVI and (c) ATR þ TVI. The tracks
are shown on a plane the passes through the middle of the central tree. The track plots are coloured using the diameter of
the droplets. Wind velocity was 3 m s¡1 at 10 m height blowing in the direction of spraying. Application rate was 500 l ha¡1
and driving speed was 1.67 m s¡1.

individual percentage drift values obtained for a particular out of the influence of the air jet before reaching the tree or
situation. This analysis was carried out for fully-leafed trees be carried further behind the trees causing environmental
which gave a maximum sedimenting drift of 20%. However, and health risks. Unfortunately, the cross-flow sprayer used
the sedimenting drift obtained from spraying bare trees could in this study, and many other commercial sprayers,
reach up to 40% (unreported tests). If the drift curves of the have little means to adjust the velocity and flow rate of air to
different setups are integrated from before the last tree to the a specific canopy. In this section, the effect of two fan
40 m point, then the ATR þ TVI and TVI configurations pro- speeds (low and high fan speed) on drift from three different
duce the same total amount of drift deposit, while the ATR nozzle arrangements fitted to a cross flow sprayer was
setup has lower total deposition drift (~16% less). This illus- analysed. The sprayer produced an airflow rate of 40,000
trates a complete analysis of the spray patterns is necessary to and 50,000 m3 h1 operating at low and high fan speeds,
correctly interpret results. respectively (Dekeyser et al., 2013). This indicates that
increasing fan speed does not necessarily lead to drastic
increases in volume airflow rates depending on the fan
3.3. Effect of fan speed on drift from three nozzle characteristics.
arrangements Figure 10 compares the sedimenting drift from three
different nozzle arrangements fitted to the crosseflow
The volumetric airflow rate and velocity of air assistance is sprayer operating at high and low fan speeds. The solid and
an important parameter that strongly influences drift from broken lines in this plot represent the percentage drift at
air assisted orchard sprayers. It transports the spray drop- high and low fan speeds, respectively. As can be seen from
lets to the target and moves the branches and leaves to Fig. 10a, b and c, the reduction in fan speed in general slightly
allow better coverage and penetration. However, depending increased the sedimenting drift close to the trees and
on the type of the canopy and the air flow, the spray can fall

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
10 b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4

Fig. 8 e Contour plots showing the time-integrated spray deposition on the ground for the three nozzle arrangements fitted
to a cross-flow sprayer: (a) ATR, (b) TVI and (c) ATR þ TVI. Wind velocity was 3 m s¡1 measured at 10 m height blowing in
the direction of spraying. Application rate was 500 l ha¡1 and driving speed was 1.67 m s¡1.

Fig. 9 e The drift curves predicted by the CFD model for an apple classical training system sprayed with the cross-flow
sprayer using three different nozzle arrangements: ATR, TVI and TVI þ ATR. The simulations were done for the
same wind velocity of 3 m s¡1 measured at 10 m height blowing in the direction of spraying. Application rate was 500 l ha¡1
and driving speed was 1.67 m s¡1.

decreased the drift values further behind the tree. This is condition considered in this study. The effect of a reduction
expected as the reduction in fan speed decreases the in fan speed can be significant when there is a strong cross
strength of the air assistance to the spray droplets causing a flow wind blowing perpendicular to the spraying direction
portion of the spray that was supposed to travel behind the (Cross, Walklate, Murray, & Richardson, 2003) or when the
tree deposit nearby. However, the reduction in the drifting canopy is very dense. The extent of drift reduction would be
distance was not significant to the specific canopy and wind much more significant when the spraying is done on very

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4 11

Fig. 10 e Effect of fan speed on the percentage drift from an apple classical training system sprayed with the cross-flow sprayer
using three nozzle arrangements: (a) ATR, (b) TVI and (c) TVI þ ATR. The solid and broken lines represent the percentage drift
obtained at high and low fan speeds, respectively. The simulations were done for the same wind velocity of 3 m s¡1 measured
at 10 m height blowing in the direction of spraying. Application rate was 500 l ha¡1 and driving speed was 1.67 m s¡1.

dense canopy or in the presence of strong cross flow winds proportion of fine droplets in the droplet size spectra increases
blowing perpendicular to the spraying direction. the spray coverage at the expense of high drift. A large pro-
portion of coarse droplets on the other hand reduces spray drift
but results in poor coverage due to the tendency of coarse
4. Discussions droplets to rebound from the leaf surface. It is possible to
reduce spray drift from orchard sprayers without compro-
In general, further drift reduction than that which could be mising the biological efficacy if it is possible to generate coarse
obtained using drift reducing nozzles can be achieved by droplets that have lower tendency to rebound. As suggested in
operating the fans at low speed as previously reported by previous research, this can be achieved by using either adju-
other researchers (Balsari et al., 2014; Landers, 2011). This was vants (Miller, Hewitt, & Bagle, 2001; Oliveira, Antuniassi, Mota,
also seen in the results presented here although the difference & Chechetto, 2013; Salyani & Cromwell, 1993; Spanoghe, De
was not significant for the specific spraying conditions Schampheleire, van der Meeren, Steurbaut, 2007) or air-
considered. It has also been shown that the number of drop- induction (drift-reducing) nozzles (Behmer, Di Prinzio,
lets which deposit on the ground close to the trees when Striebeck, & Magdalena, 2010; Derksen, Fox, Brazee, & Krause,
operating at low fan speeds could be decreased by using a 2007; Mcartney & Obermiller, 2008; Wenneker, Heijne, & Van
combination of the standard Albuz ATR orange nozzles and de Zande, 2005; Wenneker & Van de Zande, 2008; Zhu, Guler,
Albuz TVI 8002 yellow drift reducing nozzles rather than using Derksen, & Ozkan, 2005). However, some researchers reported
only Albuz TVI 8002 drift reducing nozzles. Care should no pronounced effect on droplet size from adjuvants (Fritz,
however be taken when there is a strong cross-flow wind Hoffmann, & Bagley, 2012) and similar drift profiles from all
blowing perpendicular to the spraying direction which makes conventional adjuvants (Butler Ellis & Tuck, 1999). The use of
the spray droplets highly susceptible to drift. drift reducing nozzles to reduce spray drift by generating coarse
Droplet size distribution plays an important role in deciding droplets has attracted the attention of many researchers due to
both the quality of treatment and amount of pesticide drift the ballistic behaviour of the droplets. The air-filled droplets
from air assisted orchard sprayers. The presence of a large produced by these nozzles can disintegrate into fine spray

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
12 b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4

when they impact with a solid surface and spread onto the that combining the standard Albuz ATR and TVI drift reducing
target rather than rebound. Comparison of drift from drift nozzles decreased the spray deposition close to the trees while
reducing and conventional hydraulic nozzles by some re- achieving 50% reduction in drift distance. The combined use of
searchers (Behmer et al., 2010; Wenneker et al., 2005; Zhu et al., these two nozzle types requires no complex sprayer design
2005) showed a reduction in drift by drift reducing nozzles modification, which makes it easy and cheap to implement.
which is similar to the results obtained from this CFD analysis. The presence of the standard nozzles that the farmers are
Some researchers also reported that drift reducing nozzles give familiar with may also avoid scepticism and make it easier to
higher ground deposition near the orchard boundaries (Heijne, convince them to accept change.
Wenneker, Van de Zande, & Western, 2002; Wenneker et al., The effect of high and low fan speeds on the percentage
2005; Wenneker & Van de Zande, 2008; Zhu et al., 2005) which drift from three nozzle arrangements was also analysed using
again conforms with the results of the CFD simulations per- the CFD model. The results obtained showed that fan speed
formed in this study. does not have a significant effect on drift for the spray con-
Several countries have developed their own guidelines and ditions considered in this study. The developed model can be
mitigation measures to reduce pesticides drift and most of used to do further analysis on other mitigation strategies
them included the use of drift reducing nozzles as one. The (more nozzle combinations and other sprayer designs). It can
Belgian Federal Public Service for Health, Food Chain Safety also be used to investigate the effect of other parameters (e.g.
and Environment which is responsible for registering (li- different wind speeds and directions) on drift from air-
cencing) pesticides for sale and use in Belgium has imposed assisted orchard sprayers.
eight drift mitigation requirements on pesticide product labels Finally, the presented model overcomes the following chal-
(Anon, 2004). The buffer zone requirement in this mitigation lenges of orchard drift modelling in a physically resolved way:
measures can be reduced using drift reducing nozzles. How-
ever, drift reducing nozzles should be used in combination (1) the moving spray interacts with the canopy before
with other techniques to avoid the high ground deposition reaching the drift area: by modelling the tree architec-
near the orchard boundary which impairs the reduction in ture and leaf cover, the wind velocity changes across
drift. Wenneker and Van de Zande (2008) reported on the use the tree row and droplets are captured and deviated
of shielded sprayer to overcome the problem of high deposi- before reaching the drift zone.
tion near the orchard boundaries by these nozzles. In this (2) the vertical wind profile changes from the orchard to
work, the combined use of Albuz TVI drift reducing and Albuz the neighbouring field that has a different vegetation:
ATR standard nozzles (upper three TVI and lower five ATR the airflow field is solved by means of the governing
nozzles) on a cross-flow sprayer as a way of reducing the high equations continuously across the trees into and over
ground deposition found near the orchard boundaries was the drift zone. From a canopy profile inside the orchard,
explored. It was interesting to see that it is possible to reduce the flow develops into an atmospheric boundary profile
the near orchard ground deposition while maintaining the across the neighbouring drift field with a specified
same drift reduction as with drift reducing nozzles. In this roughness height of the grass field. Flow paths around
work, simulations were performed for only one combination the specific tree training system are resolved and affect
of nozzles and one sprayer type. It would be interesting to the drift profile.
extend investigations to other combinations of these two (3) the moving air jet from the air assistance cannot be
nozzle types and also to other sprayer designs. The results ignored because the magnitude of the air jet velocity is
reported in this work also showed that it is possible to reduce typically higher than the wind speed: a dynamic model
drift further by using the drift reducing nozzles on the cross- is implemented that resolves the moving sprayer outlet
flow sprayer at a low fan speed. The high ground deposition of the orchard sprayer over the tree row.
near the orchard boundary while operating the sprayer at low
fan speed can be mitigated by using a combination of drift
reducing TVI and standard ATR orange nozzles. Acknowledgement

The financial support of the Institute for the Promotion of


5. Conclusions Innovation by Science and Technology in Flanders (project
IWT 080528) is gratefully appreciated.
A CFD drift model of air assisted orchard spraying was suc-
cessfully validated. This model was then used to study the ef-
fect of nozzle arrangement and fan speed on drift. One references
common disadvantage of using drift reducing nozzles as a drift
mitigation strategy is the high ground deposition observed near
the orchard boundaries. This is especially important in areas Akesson, B. N. B., & Yates, W. E. (1964). Problems relating to
where orchards are close to surface waters. This work investi- application of agricultural chemicals and resulting drift
residues. Annual Review of Entomology, 9, 285e318.
gated the potential of combining drift reducing TVI and stan-
Anon. (2004). Maatregelen ter beperking van de verontreiniging
dard Albuz ATR nozzles as a drift mitigation strategy using CFD.
van oppervlaktewater door gewasbeschermingsmiddelen. In
The results of the study showed the potential of CFD modelling Federal public service of public health, food chain security and
as a tool to investigate drift mitigation strategies. The analysis environment, DG animals, plants & food (p. 12). Brussels: Service
done using one combination of the two nozzle types showed Pesticides and Fertilizers. http://www.fytoweb.fgov.be/NL/

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4 13

doc/driftreducerende%20maatregelen%20voor% optimization of air assistance and nozzle type for orchard


20gewasbeschermingsmiddelen.pdf. spraying by CFD modelling. Aspects of Applied Biology, 122,
Arvidsson, T., Bergstro € m, L., & Kreuger, J. (2011). Spray drift as 453e458.
influenced by meteorological and technical factors. Pest Duga, A. T., Dekeyser, D., Ruysen, K., Bylemans, D., Nuyttens, D.,
Management Science, 67(5), 586e598. http://dx.doi.org/10.1002/ Nicolai, B. M., et al. (2015). Numerical analysis of the effects of
ps.2114. wind and sprayer type on spray distribution in different
Baetens, K., Ho, Q. T., Nuyttens, D., De Schampheleire, M., orchard training systems. Boundary-Layer Meteorology, 157(3),
Endalew, A. M., Hertog, M. L. A. T. M., et al. (2009). A validated 1e19. http://dx.doi.org/10.1007/s10546-015-0064-2.
2-D diffusion-advection model for prediction of drift from Duga, A., Ruysen, K., Dekeyser, D., Nuyttens, D., Bylemans, D.,
ground boom sprayers. Atmospheric Environment, 43(9), Nicolaıü, B., et al. (2015a). CFD based analysis of the effect of wind
1674e1682. http://dx.doi.org/10.1016/j.atmosenv.2008.12.047. in orchard spraying. Chemical Engineering Transactions, 44, 1e6.
Baetens, K., Nuyttens, D., Verboven, P., De Schampheleire, M., Duga, A. T., Ruysen, K., Dekeyser, D., Nuyttens, D., Bylemans, D.,
Nicolaı̈, B., & Ramon, H. (2007). Predicting drift from field Nicolai, B. M., et al. (2015b). Spray deposition profiles in pome
spraying by means of a 3D computational fluid dynamics fruit trees: Effects of sprayer design, training system and tree
model. Computers and Electronics in Agriculture, 56(2), 161e173. canopy characteristics. Crop Protection, 67, 200e213. http://
Balsari, P., Gil, E., Marucco, P., Gallart, M., Bozzer, C., Llop, J., et al. dx.doi.org/10.1016/j.cropro.2014.10.016.
(2014). Study and development of a test methodology to assess Egan, J. F., Bohnenblust, E., Goslee, S., Mortensen, D., & Tooker, J.
potential drift generated by air-assisted sprayers. Aspects of (2014). Herbicide drift can affect plant and arthropod
Applied Biology, 122, 339e346. communities. Agriculture, Ecosystems & Environment, 185,
Behmer, S., Di Prinzio, A., Striebeck, G., & Magdalena, J. (2010). 77e87. http://dx.doi.org/10.1016/j.agee.2013.12.017.
Evaluation of low-drift nozzles in agrochemical applications Endalew, A. M., Debaer, C., Rutten, N., Vercammen, J., Delele, M. A.,
in orchards. Chilean Journal of Agricultural Research, 70(3), Ramon, H., et al. (2010a). A new integrated CFD modelling
498e502. http://dx.doi.org/10.4067/S0718-58392010000300017. approach towards air-assisted orchard spraying. Part I. Model
Butler Ellis, M. C., & Tuck, C. R. (1999). How adjuvants influence development and effect of wind speed and direction on sprayer
spray formation with different hydraulic nozzles. Crop airflow. Computers and Electronics in Agriculture, 71(2), 128e136.
Protection, 18(2), 101e109. http://dx.doi.org/10.1016/S0261- http://dx.doi.org/10.1016/j.compag.2009.11.005.
2194(98)00097-0. Endalew, A. M., Debaer, C., Rutten, N., Vercammen, J.,
Butler Ellis, M. C., Lane, A. G., O'Sullivan, C. M., Miller, P. C. H., & Delele, M. A., Ramon, H., et al. (2010b). Modelling pesticide
Glass, C. R. (2010). Bystander exposure to pesticide spray drift: flow and deposition from air-assisted orchard spraying in
New data for model development and validation. Biosystems orchards: A new integrated CFD approach. Agricultural and
Engineering, 107(3), 162e168. http://dx.doi.org/10.1016/ Forest Meteorology, 150(10), 1383e1392. http://dx.doi.org/
j.biosystemseng.2010.05.017. 10.1016/j.agrformet.2010.07.001.
Cross, J. V., Walklate, P. J., Murray, R. A., & Richardson, G. M. Endalew, A. M., Debaer, C., Rutten, N., Vercammen, J., Delele, M. A.,
(2003). Spray deposits and losses in different sized apple trees Ramon, H., et al. (2011). Modeling the effect of tree foliage on
from an axial fan orchard sprayer: 3. Effects of air volumetric sprayer airflow in orchards. Boundary-Layer Meteorology, 138,
flow rate. Crop Protection, 22(2), 381e394. http://dx.doi.org/ 139e162. http://dx.doi.org/10.1007/s10546-010-9544-6.
10.1016/S0261-2194(02)00192-8. Endalew, A. M., Hertog, M., Gebrehiwot, M. G., Baelmans, M.,
De Schampheleire, M., Baetens, K., Nuyttens, D., & Spanoghe, P. Ramon, H., Nicolaı̈, B. M., et al. (2009). Modelling airflow within
(2008). Spray drift measurements to evaluate the Belgian drift model plant canopies using an integrated approach. Computers
mitigation measures in field crops. Crop Protection, 27(3e5), and Electronics in Agriculture, 66(1), 9e24.
577e589. http://dx.doi.org/10.1016/j.cropro.2007.08.017. FOCUS. (2007). Landscape and mitigation factor report of the FOCUS
Dekeyser, D., Duga, A. T., Verboven, P., Endalew, A. M., Working group on landscape and mitigation factors in ecological risk
Hendrickx, N., & Nuyttens, D. (2013). Assessment of orchard assessment. EC Document Reference SANCO/10422/2005 V.2.0,
sprayers using laboratory experiments and computational fluid Aquatic Risk Assessment. Extended Summary and
dynamics modelling. Biosystems Engineering, 114(2), 157e169. Recommendations.
http://dx.doi.org/10.1016/j.biosystemseng.2012.11.013. Fritz, B. K., Hoffmann, W. C., & Bagley, W. E. (2012). Effects of
Delele, M. A., De Moor, A., Sonck, B., Ramon, H., Nicolaı̈, B. M., & formulated glyphosate and adjuvant tank mixes on
Verboven, P. (2005). Modelling and validation of the air flow atomization from aerial application flat fan nozzles. Pesticide
generated by a cross flow air sprayer as affected by travel Formulation and Delivery Systems: Innovating Legacy Products for
speed and fan speed. Biosystems Engineering, 92(2), 165e174. New Uses, 32, 80e95. http://dx.doi.org/10.1520/STP104451.
Delele, M. A., Jaeken, P., Debaer, C., Baetens, K., Endalew, A. M., Graham, D. I., & Moyeed, R. A. (2002). How many particles for my
Ramon, H., et al. (2007). CFD prototyping of an air-assisted Lagrangian simulations? Powder Technology, 125(2e3), 179e186.
orchard sprayer aimed at drift reduction. Computers and http://dx.doi.org/10.1016/S0032-5910(01)00504-6.
Electronics in Agriculture, 55(1), 16e27. Gregorio, E., Rocadenbosch, F., Sanz, R., & Rosell-Polo, J. (2015).
Derksen, R. C., Fox, R. D., Brazee, R. D., & Krause, C. R. (2007). Eye-safe lidar system for pesticide spray drift measurement.
Coverage and drift produced by air induction and conventional Sensors, 15(2), 3650e3670. http://dx.doi.org/10.3390/
hydraulic nozzles used for orchard applications. In Transactions s150203650.
of the ASABE (pp. 1497e1501). American Society of Agricultural Gregorio, E., Rosell-Polo, J. R., Sanz, R., Rocadenbosch, F.,
Engineers. http://dx.doi.org/10.13031/2013.23941. Solanelles, F., Garcera  , C., et al. (2014). LIDAR as an alternative
Donkersley, P., & Nuyttens, D. (2011). A meta-analysis of spray to passive collectors to measure pesticide spray drift.
drift sampling. Crop Protection, 30(7), 931e936. http:// Atmospheric Environment, 82, 83e93. http://dx.doi.org/10.1016/
dx.doi.org/10.1016/j.cropro.2011.03.020. j.atmosenv.2013.09.028.
Dorr, G. J., Hewitt, A. J., Adkins, S. W., Hanan, J., Zhang, H., & Heijne, B., Wenneker, M., Van de Zande, J. C., & Western, N. M.
Noller, B. (2013). A comparison of initial spray characteristics (2002). Air inclusion nozzles don't reduce pollution of surface
produced by agricultural nozzles. Crop Protection, 53, 109e117. water during orchard spraying in The Netherlands. Aspects of
http://dx.doi.org/10.1016/j.cropro.2013.06.017. Applied Biology, 193e199.
Duga, A. T., Defraeye, T., Nicolai, B., Dekeyser, D., Nuyttens, D., Hilz, E., & Vermeer, A. W. P. (2013). Spray drift review: The extent
Bylemans, D., et al. (2014). Training system dependent to which a formulation can contribute to spray drift reduction.

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010
14 b i o s y s t e m s e n g i n e e r i n g x x x ( 2 0 1 6 ) 1 e1 4

Crop Protection, 44, 75e83. http://dx.doi.org/10.1016/ potential by means of different wind tunnel evaluation
j.cropro.2012.10.020. methods. Biosystems Engineering, 103(3), 271e280. http://
Holterman, H., & Van de Zande, J. (2008). The cascade drift dx.doi.org/10.1016/j.biosystemseng.2009.04.001.
module: A GIS-based study on regional pesticide deposition. Oliveira, R. B. De, Antuniassi, U. R., Mota, A. A. B., &
Aspects of Applied Biology, 84, 83e90. Chechetto, R. G. (2013). Potential of adjuvants to reduce drift in
Holterman, & Van de Zande. (2010). The cascade drift model: First agricultural spraying. Engenharia Agrı́cola, 33(5), 986e992.
results of a realistic study on regional pesticide deposition. http://dx.doi.org/10.1590/S0100-69162013000500010.
International Advances in Pesticide Application, Aspects of Applied Raupach, M. R., Woods, N., Dorr, G., Leys, J. F., & Cleugh, H. A.
Biology, 99, 367e374. (2001). The entrapment of particles by windbreaks.
Holterman, H., Van de Zande, J., Porskamp, H. A. J., & Atmospheric Environment, 35(20), 3373e3383. http://dx.doi.org/
Huijsmans, J. (1997). Modelling spray drift from boom 10.1016/S1352-2310(01)00139-X.
sprayers. Computers and Electronics in Agriculture, 19(1), 1e22. Rautmann, D., Streloke, M., & Winkler, R. (2001). New basic drift
http://dx.doi.org/10.1016/S0168-1699(97)00018-5. values in the authorization procedure for plant protection
ISO 22866. (2005). International Standard: Crop protection equipment e products. Mitt. Biol. Bundesanst. Land Fiorst-Wirtsch, 383,
Methods for field measurement of spray drift. 133e141.
Khot, L. R., Miller, D. R., Hiscox, A. L., Salyani, M., Walker, T. W., & Salyani, M., & Cromwell, R. P. (1993). Adjuvants to reduce drift
Farooq, M. (2011). Extrapolation of droplet catch from handgun spray application. ASTM STP 1146 Pesticide
measurements in aerosol application treatments. Atomization Formulations and Application Systems, 12, 363e376.
and Sprays, 21(2), 149e158. http://dx.doi.org/10.1615/ Salyani, M., Miller, D. R., Farooq, M., & Sweeb, R. D. (2013). Effects
AtomizSpr.2011002846. of sprayer operating parameters on airborne drift from citrus
Kruckeberg, J. P., Hanna, H. M., Steward, B. L., & Darr, M. J. (2012). air-carrier sprayers. Agricultural Engineering International: CIGR
The relative accuracy of DRIFTSIM when used as a real-time Journal, 15(1), 27e36.
spray drift predictor. Transactions of the ASABE, 55(4), 1159e1165. Spanoghe, P., Schampheleire, M. De, & Meeren, P. Van Der (2007).
Kuzmin, D., Mierka, O., & Turek, S. (2007). On the implementation Influence of agricultural adjuvants on droplet spectra. Pest
of the kε turbulence model in incompressible flow solvers Management, 63, 4e16. http://dx.doi.org/10.1002/ps.
based on a finite element discretisation. International Journal of Stoughton, T. E., Miller, D. R., Yang, X., & Ducharme, K. M. (1997).
Computing Science and Mathematics, 1(2/3/4), 193. http:// A comparison of spray drift predictions to lidar data.
dx.doi.org/10.1504/IJCSM.2007.016531. Agricultural and Forest Meteorology, 88(1e4), 15e26. http://
Landers, A. (2011). Improving spray deposition with engineering dx.doi.org/10.1016/S0168-1923(97)00056-7.
innovation e What a difference a decade makes. Research Teske, M. E., Bird, S. L., Esterly, D. M., Curbishley, T. B., Ray, S. L., &
Focus, 1e7. Perry, S. G. (2002). AgDRIFT: A model for estimating near-field
Lazzaro, L., Otto, S., & Zanin, G. (2008). Role of hedgerows in spray drift from aerial applications. Environmental Toxicology
intercepting spray drift: Evaluation and modelling of the and Chemistry/SETAC, 21(3), 659e671. http://dx.doi.org/10.1002/
effects. Agriculture, Ecosystems & Environment, 123(4), 317e327. etc.5620210327.
http://dx.doi.org/10.1016/j.agee.2007.07.009. Teske, M. E., Thistle, H. W., & Ice, G. G. (2003). Technical advances
Lee, I. B., Bitog, J. P. P., Hong, S. W., Seo, I. H., Kwon, K. S., in modelling aerially applied sprays. Transactions of the ASAE,
Bartzanas, T., et al. (2013). The past, present and future of CFD 46(4), 985e996.
for agro-environmental applications. Computers and Electronics Van De Zande, J. C., Huijsmans, J. F. M., Porskamp, H. A. J.,
in Agriculture, 93, 168e183. http://dx.doi.org/10.1016/ Michielsen, J. M. G. P., Stallinga, H., Holterman, H. J., et al.
j.compag.2012.09.006. (2008). Spray techniques: How to optimise spray deposition
Mcartney, S. J., & Obermiller, J. D. (2008). Comparative and minimise spray drift. Environmentalist, 28(1), 9e17. http://
performance of air-induction and conventional nozzles on an dx.doi.org/10.1007/s10669-007-9036-5.
axial fan sprayer in medium density apple orchards. Wenneker, M., Heijne, B., & Van de Zande, J. C. (2005). Effect of
HortTechnology, 18(3), 365e371. air induction nozzle (coarse droplet), air assistance and one-
Miller, P. C. H., Hewitt, A. J., & Bagle, W. E. (2001). Adjuvant effects sided spraying of the outer tree row on spray drift in orchard
on spray characteristics and drift potential. ASTM STP 1414 spraying. Annual Review of Agricultural Engineering, 4(1),
Pestesticide Formulations and Application Systems, 21, 175e184. 116e128.
Miller, D. R., Saliyani, M., & Hiscox, A. B. (2003). Remote Wenneker, M., & Van de Zande, J. (2008). Drift reduction in
measurement of spray drift from orchard sprayers using orchard spraying using a cross flow sprayer equipped with
LIDAR. Transactions of the ASABE (p. Paper number 031093). reflection shields (Wanner) and air injection nozzles.
Nsibande, S. A., Dabrowski, J. M., van der Walt, E., Venter, A., & Agricultural Engineering International: The CIGR Journal.
Forbes, P. B. C. (2015). Validation of the AGDISP model for Manuscript ALNARP, 8, 014.
predicting airborne atrazine spray drift: A South African Wilson, N. R., & Shaw, R. H. (1977). A higher order closure model
ground application case study. Chemosphere, 138, 454e461. for canopy flow.pdf. Journal of Applied Meteorology, 16,
Nuyttens, D., Baetens, K., De Schampheleire, M., & Sonck, B. 1197e1205.
(2007a). Effect of nozzle type, size and pressure on spray Zhu, H., Guler, H., Derksen, R. C., & Ozkan, H. E. (2005).
droplet characteristics. Biosystems Engineering, 97, 333e345. Comparison of airborne and ground spray deposits with
http://dx.doi.org/10.1016/j.biosystemseng.2007.03.001. hollow cone nozzle, low drift nozzle and drift retardant. In 9th
Nuyttens, D., De Schampheleire, M., Baetens, K., & Sonck, B. International Congress on Mechanization and Energy in Agriculture
(2007b). The influence of operator-controlled variables on & 27th International Conference of CIGR Section IV: The Efficient Use
spray drift from field crop sprayers. Transactions of the ASABE, of Electricity and Renewable Energy Sources in Agriculture (pp.
50(4), 1129e1140. http://dx.doi.org/10.13031/2013.23622. 27e29) (Izmir-Turkey).
Nuyttens, D., Taylor, W. A., De Schampheleire, M., Verboven, P., &
Dekeyser, D. (2009). Influence of nozzle type and size on drift

Please cite this article in press as: Duga, A. T., et al., Development and validation of a 3D CFD model of drift and its application to air-
assisted orchard sprayers, Biosystems Engineering (2016), http://dx.doi.org/10.1016/j.biosystemseng.2016.10.010

You might also like