You are on page 1of 13

Computers and Electronics in Agriculture 76 (2011) 218230

Contents lists available at ScienceDirect

Computers and Electronics in Agriculture


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

Original papers

Aerodynamic analysis and CFD simulation of several cellulose evaporative


cooling pads used in Mediterranean greenhouses
b , A.M. Prez a
A. Franco a , D.L. Valera b, , A. Pena
a
b

ETSIA, Universidad de Sevilla, Ctra. Utrera km. 1, 41013 Sevilla, Spain


Universidad de Almera, Ctra. Sacramento s/n, 04120 Almera, Spain

a r t i c l e

i n f o

Article history:
Received 27 July 2010
Received in revised form 14 January 2011
Accepted 30 January 2011
Keywords:
Greenhouse
Aerodynamic analysis
CFD
Evaporative cooling
Fan and pad
Pressure drop

a b s t r a c t
The present work makes an aerodynamic analysis and computational uid dynamics (CFD) simulation of
the four commercial models of corrugated cellulose evaporative cooling pads that are most widely used
in Mediterranean greenhouses. The geometric characteristics of the pads have been determined as well as
the volume of water they retain at different ows of water, thus obtaining the mean thickness of the sheet
of water which runs down them and their porosity. By means of low velocity wind tunnel experiments,
the pressure drop produced by the pads has been recorded at different wind speeds and water ows. In
this way it has been possible to obtain the relationship of the permeability and the inertial factor with pad
porosity using a cubic type equation. Finally, a CFD simulation with a 3D model has been carried out for
both dry pads (Qw = 0 l s1 m2 ) and wet ones (Qw = 0.256 l s1 m2 ), nding good correlation between the
simulated and experimental pressure drop, with maximum differences of 9.08% for dry pads and 15.53%
for wet ones at an airspeed of 3 m s1 .
2011 Elsevier B.V. All rights reserved.

1. Introduction
The surface area of greenhouses is increasing worldwide.
Approximately 20% of this area is concentrated in the Mediterranean basin, consisting for the most part of only rudimentary
greenhouses (Baille, 2001). The Spanish Mediterranean coast has
45,000 ha of greenhouses (Castilla and Hernndez, 2005) and the
southeast of the country, particularly the province of Almera, is
the location of the highest concentration of greenhouses in the
world, with 30,000 ha of farms, representing 4% of the global total
(Molina-Aiz, 2010).
The protected horticulture sector in southern Europe is currently facing strong international competition, mainly from areas
such as the north of Africa, where production costs, including labour
costs, are substantially lower. In order to improve the sustainability of greenhouse crops it is necessary to improve the nal
quality of production, increase crop yield and modify the periods
of maximum production. These requirements have led to constant
advances in greenhouse technology.
As a result of this technication, evaporative cooling systems
are being implanted in areas with high spring-summer temperatures such as the Mediterranean basin. These systems reduce the
temperature inside the greenhouse by increasing the air humidity

Corresponding author. Tel.: +34 950015546; fax: +34 950015491.


E-mail addresses: afranco@us.es (A. Franco), dvalera@ual.es (D.L. Valera),

apfernan@ual.es (A. Pena),


tao@us.es (A.M. Prez).
0168-1699/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.compag.2011.01.019

content. Thus, by maintaining suitable hygrometric levels in the


development stages of crops with little foliage and high evapotranspiration requirements, the transplant date of some autumn-winter
crops can be brought forward to mid-August. As a result, the crop
can be harvested at a time when the produce fetches higher prices.
Evaporative refrigeration can be carried out by directly spraying
water inside the greenhouse and combining it with natural ventilation (Fog/Mist Systems) or by obliging the outside air to pass into
the greenhouse through moistened evaporative pads combined
with mechanical ventilation (FanPad Cooling Systems).
Fan and pad cooling systems constitute a substantial improvement in climate control for greenhouses (Sethi and Sharma, 2007a)
and poultry houses (Dagtekin et al., 2009). An evaporative pad is
a permeable screen which is saturated with water by means of an
irrigation system on its upper part. The temperature of the air passing through the screen drops, and these air then cools the inside air
on mixing with it.
The pads are located along part or all of the side of the greenhouse. They are continuously moistened by an irrigation and
recirculation system. Powerful extractor fans installed on the opposite side of the greenhouse create the necessary suction to ensure
that outside air passes into the greenhouse through the pads. When
the cool damp air from the pad comes into contact with the mass
of warm dry air inside the greenhouse, the air velocity falls, which
brings about a sharp increase in the turbulence of the air ow
(turbulence intensity, turbulence kinetic energy, turbulence kinetic
energy dissipation rate and slope of the spectrum of energy density). This increase in turbulence is less when there is a crop in the

A. Franco et al. / Computers and Electronics in Agriculture 76 (2011) 218230

Nomenclature
A
As
e
K
Kf
l
le
le /l
Qw
Re
V
Va
Vl
Vs
VT
%diff
P

surface area per unit volume (m2 m3 )


surface area of pad media (m2 )
mean thickness of the sheet of water (m)
permeability (m2 )
Forchheimers permeability (m2 )
pad thickness (m)
characteristic length of pad (m)
characteristic parameter of pad geometry (adimensional)
water ow applied (l s1 m2 )
Reynolds number (adimensional)
airspeed (m s1 )
volume of air in the pad (m3 )
retained water volume (m3 )
solid volume of the cellulose sheets (m3 )
total sample volume of pad (m3 )
percentage difference (%)
pressure drop (Pa)

Greek letters

factor inertial de Forchheimer (adimensional)



cubic law inertial factor (adimensional)

uid density (kg m3 )

dynamic viscosity (Pa s)

porosity (m3 m3 )

greenhouse, which indicates that the crop has a dispersion effect on


the air momentum, which is characteristic of porous media (Lopez
et al., 2010). As the air moves from one end of the greenhouse to
the other, undesirable gradients of temperature, humidity and CO2
develop throughout the greenhouse; the largest are to be expected
at around midday when the intensity of solar radiation is greatest (Teitel et al., 2010), and in circumstances of low leaf area index
and low drag coefcient. This system requires extremely air-tight
structures to ensure that all the air entering the greenhouse does
so through the pads. This fact has meant that in structures that
are not so hermetic, such as Almera-type greenhouses, another
system has been more widely applied, namely the fog system. Nevertheless, the saturation efciency of the padfan system is greater
(Katsoulas et al., 2009), it is cheaper to install and consumes less
water and energy (Sethi and Sharma, 2007b) than the fog system.
Moreover, the different types of greenhouses used in southeast
Spain are evolving towards more hermetic multi-tunnel structures,
which have increased from 0.6% of the total greenhouse area in the
province of Almera in 1997, to 2.5% in 2006 (Valera and MolinaAiz, 2008). This trend is continuing at present (Molina-Aiz, 2010)
and the increasing number of multi-tunnel greenhouses on the
Mediterranean coast is allowing more and more padfan systems
to be installed.
Greenhouse cooling by means of the padfan system obviously
involves higher energy and water consumption than the use of the
traditional system of natural ventilation and transpiration of the
crop. However, this increase in production costs can be offset by
the corresponding increase in earliness, quality and yield of the
crop, allowing growth in arid zones during the summer period.
This allows growers to produce over a longer period and/or to commence the following season earlier, thus modifying the periods of
maximum production.
The water consumption of the pads increases linearly with the
ventilation rate (Sabeh et al., 2007), registering values of between
3.2 and 10.3 l m2 of greenhouse for 150 mm cellulose pads in semiarid conditions. In conditions of extreme aridity the recorded values

219

were between 7.9 and 16.3 l m2 of greenhouse for 100 mm corrugated cellulose pads (Al-Helal, 2007). For 100 mm pads in the
province of Almera (Spain) values were recorded of 146.3 l d1 m2
of evaporative pad (Franco et al., 2010), with an air velocity through
the pad of 1.27 m s1 (ANSI/ASABE Standards, 2008), producing a
mean reduction of 8 C compared to the outside temperature (Sethi
and Sharma, 2007a) working 8 h d1 on average. However, the crop
requires less irrigation (Montero, 2006), as its transpiration rate
falls by 31%. Therefore, the total increase in water requirement
of a greenhouse tted with a padfan cooling system is only 19%
(Katsoulas et al., 2009). Moreover, if this system is combined with
shading, the thermal gradient decreases (Kittas et al., 2003) and
electrical consumption is reduced by around 8% (Willits and Peet,
2000).
The vast majority of new greenhouses built in the Mediterranean area incorporate padfan cooling systems using corrugated
cellulose pads. These are more costly than alternative local materials but they are a rigid porous medium that only requires a simple
support structure, they are easy to maintain, do not emit particles
and are durable. The present study therefore focuses on this type
of pad, analyzing the most widely used commercial models.
Rising energy costs, together with scant water resources in most
areas of intensive production, urge the use of evaporative cooling
systems that are economical and highly water and energy efcient.
Choosing the most suitable evaporative pad requires knowledge
of its different parameters. ANSI/ASABE Standards (2008) recommends values of airspeed through the pads, minimum water ow
to be applied and other parameters for only two types of pad and
two different thicknesses: Aspen ber pads of 50 and 100 mm, and
corrugated cellulose pads of 100 and 150 mm.
Due to the intricate geometry of corrugated cellulose pads
and the complex processes of heat and mass transfer that occur
inside them, semi-empirical methods are still used for their design
(Beshkani and Hosseini, 2006).
Evaporative pads make it more difcult for outside air to ow
into the greenhouse. Expressed as pressure drop, this resistance
to air ow depends on the geometric characteristics of the pad,
on the amount of water applied and on the air ow (Franco et al.,
2010). Less resistance to air ow means lower energy costs, as the
fans require less energy to maintain the ventilation rate. Finding a
procedure to simulate new pad designs that reduce resistance to
the ow of air and therefore increase their efciency is a key factor
for reducing costs.
The aerodynamic characteristics of the different types of cellulose evaporative pads at different airspeeds and water ows
must be known in order to analyze their pressure drop. Valera
et al. (2006) carried out assays with insect-proof screens using two
devices that suction air through the screens, obtaining data of pressure drop as a function of airspeed. This type of analysis allows us
to obtain simple ratios between permeability and the inertial factor
as a single function of porosity.
On the other hand, in recent years computational uid dynamics
(CFD) has proved to be a most useful tool for simulating the interaction of liquids and gases with complex surfaces, and the many
works published on the subject have achieved results that are very
close to real results. Verication of the data obtained by CFD is usually carried out in a wind tunnel, in other scale models or by means
of direct measures.
According to Molina-Aiz et al. (2010), over the last 25 years,
computational uid dynamics (CFD) has been increasingly used to
describe quantitatively and qualitatively the natural ventilation in
greenhouses (Mistriotis et al., 1997; Boulard et al., 1999; MolinaAiz et al., 2004; Campen, 2006; Lee et al., 2006; Baeza et al., 2006;
Tong et al., 2009; Teitel et al., 2008) and to model other systems
of climatic control such as thermal screens (Montero et al., 2005),
fan ventilation (Fidaros et al., 2008) or fog systems (Kim et al.,

220

A. Franco et al. / Computers and Electronics in Agriculture 76 (2011) 218230

2008). CFD has also been used to study the distribution and deposition of fungal spores in greenhouses (Roy et al., 2006) or pesticide
dispersion (Bartzanas et al., 2006). At present, simulations of CFD
have achieved a high level of complexity, including the radiation
exchanges between the atmosphere and the greenhouse surfaces
(Ould Khaoua et al., 2006; Bournet et al., 2007) and the transfer of
heat and water vapour between the crop and the air (Boulard et al.,
2002; Roy et al., 2008; Majdoubi et al., 2009; Fatnassi et al., 2009).
Simulation of evaporative pads using CFD provides the opportunity to improve their performance by increasing their cooling
efciency and reducing resistance to the airow through the pad,
which results in energy saving and therefore a reduction in environmental impact.
The present work therefore has the following three aims: (i)
to determine experimentally the inuence of water ow regimes
applied to evaporative pads on their porosity and on the pressure
drop they produce when air passes through them, (ii) to characterize the airow on passing through the evaporative pads by means of
permeability and the inertial factor coefcient of cubic law, and (iii)
to create a numeric model using CFD, and to validate these results
with experimental data of the inuence of air ow and volume
of water on the pressure drop for the four commercial corrugated
cellulose evaporative pads that are available on the market.

2. Materials and methods


2.1. Geometric characterization of the evaporative pads
The pads are made up of sheets of corrugated cellulose that are
stuck together alternating the angles of incidence on the horizontal
so that they do not coincide (Fig. 1). In this way we achieve greater
transfer surface (m2 m3 ), greater mechanical resistance and lower
resistance to the passage of air and water. The length and width
of the undulation of the sheets, together with the angles of incidence and the thickness of the pad are considered the characteristic
geometric parameters.
Four commercial models of evaporative pads made by two different manufacturers were tested. These are the most commonly
used models in the Mediterranean region. The manufacturer G&R
(Gigola and Riccardi, Italy) commercializes a 100 mm thick model
with angles of 4545 , while Munters (Kista, Sweden) markets
three models, two with angles of 6030 that are 50 and 100 mm
thick, and one with angles of 4545 that is 100 mm thick, but with
length and width of undulations different from the G&R model.
Three different samples of each type were tested in the wind tunnel. These samples were 0.60 m wide by 0.65 m high, i.e., more than
the minimum height of 0.60 m as recommended by ANSI and ASABE
(ANSI/ASABE Standards, 2008).
A detailed study was made of the geometric characteristics of
the cellulose pads tested. The geometric parameters are as follows:
angles of incidence of the sheets that make up the pads ( ), thickness of the pad (mm), number of sheets per meter of pad width
(ud, m1 ), thickness of the sheets (mm), undulation length (mm),
undulation width (mm), specic area of the pad (m2 m3 ), porosity (m3 m3 ), and a non-dimensional geometric parameter (le /l),
where le is the characteristic length (m) and l is the thickness of the
panel (m). The characteristic length (m) is dened as le = V/As = A1 ,
where V is the volume occupied by the porous medium (m3 ), As
is the area of transfer of the pad (m2 ), and A is the specic area
of the pad or the area per unit of pad volume (m2 m3 ). For the
length and width of the undulation of the sheets, the caliper and
thickness were measured with a micrometer. To calculate the specic area, we used image software (Matrox Inspector v.2.2, Matrox
Electronic Systems Ltd., Quebec, Canada). Finally, the dry porosity
was calculated as one minus the ratio between the solid volume

and the total volume of the pad. Table 1 shows the geometric characteristics obtained in our study. This geometry is fundamental to
carry out the uid domain for the aerodynamic analysis of the pads
and the CFD simulations of the evaporative pads.
Prior the tests, the pads were immersed in water during 24 h and
that all the owing water remains on the pads surface and is not
retained inside the pads solid matrix. The volume of water retained
by the pad must be determined as a function of the ow of water
applied to calculate the pad porosity and to simulate its behavior
when wet. In order to determine the geometry of the uid domain
we have made a simplication by considering that the water displaced by the pad due to gravity maintaining constant the water
sheet thickness over the whole transfer surface. Said mean thickness of the sheet of water is easy to calculate given the retained
water volume and the transfer surface of the sample (Eq. (1)). Prior
the tests, the pads were immersed in water during 24 h and that all
the owing water remains on the pads surface and is not retained
inside the pads solid matrix. The sheet of water reduces the porosity of the evaporative pad, and consequently increases its resistance
to the airow.
e=

Vl
Vl
=
As
AVT

(1)

where e is the mean thickness of the sheet of water (m), Vl the volume of water retained by the pad (m3 ), As the pads area of transfer
(m2 ), A the pads specic area (m2 m3 ) and VT the total volume of
the pad sample (m3 ).
The porosity () of the different pads under different working
conditions can be correlated to the geometry (le /l) and the water
ow applied (Qw ), the same as the pressure loss (Milosavljevic and
Heikkil, 2001; Franco et al., 2010), as follows:
=k

 l a
e

b
(1 + Qw
)

(2)

Carrying out a nonlinear regression analysis we can obtain the values of the parameters k, a and b.
2.2. Assay equipment and procedures
In order to compare and validate the results of the CFD simulations regarding the inuence of air and water ows and pressure
drop with the experimental data, a low-speed open-circuit wind
tunnel was used with a circular cross-section of 38.8 cm diameter.
The wind tunnel was designed and constructed in the Department
of Rural Engineering of the University of Almera (Valera et al.,
2006). A uniform and stable air ow was achieved (as reported by
Fang et al., 2001) under controlled conditions of temperature and
humidity.
To carry out tests with evaporative pads in the wind tunnel, a
specic test frame was designed to incorporate the pads (Franco
et al., 2010). This frame consisted of a galvanized metal structure
with a water distribution system incorporated into the top part
(Fig. 2). The water distribution system was constructed of a 20 mm
diameter PVC pipe with 2 mm holes 65 mm apart. In the lower part
of the frame, a water collection system allowed water to drain by
gravity into a tank, before being recycled by a 12 V axial pump.
Water ow at the entrance was controlled by varying the voltage
of the continuous-current hydraulic pump and readings from the
rotameter (owmeter) with an average range of 322 l min1 and
an error of 4%.
The geometry of the frame was suitable to ensure that the assays
with moistened pads were as similar as possible to conditions in
the greenhouse. Moreover, the rectangular chassis was necessary
to hold the rectangular sample of the pad, the system for applying water and distributing it uniformly and the drainage system.
The pad sample assayed must be rectangular in order to ensure

A. Franco et al. / Computers and Electronics in Agriculture 76 (2011) 218230

221

Fig. 1. Composition of the four evaporative pads analyzed: a) 4545 100 mm by G&R; b) 4545 100 mm by Munters; c) 6030 100 mm by Munters; d) 6030 50 mm by
Munters.
Table 1
Characteristics of the cellulose pads assayed.
Manufacturer

G&R
Munters
Munters
Munters

Pad thickness
(mm)
100
100
100
50

Angles ( )
(units/m)

No. sheets

4545
4545
6030
6030

157
142
132
208

Thickness of
sheets (mm)

Length of
undulation
(mm)

Width of
undulation
(mm)

0.2130.213
0.2280.228
0.1920.191
0.2220.219

19.519.5
20.520.5
18.517
1211

6.376.37
7.007.00
7.507.50
4.804.80

Fig. 2. Diagram of the water recirculation system for moistening the evaporative
pad, the system for measuring the water ow applied and the weighing system.

Specic area
(m2 m3 )

le /l

Dry porosity
(m3 m3 )

391.114
347.114
361.516
556.752

2.557 102
2.888 102
2.766 102
3.592 102

0.957
0.959
0.965
0.937

uniform distribution of the water over the whole surface of the


cellulose sheets it comprises. The chassis was completely airtight,
only allowing the passage of air through the circular section of
the chassis which coincided with the circular section of the wind
tunnel.
Three different samples were tested for each pad model. These
samples were 0.60 m wide by 0.65 m high. The pads were immersed
in water for 24 h before each test in order for them to be totally
saturated. Four different water ow rates were tested, both above
and below the 6.2 l min1 m1 minimum recommended by ANSI
and ASABE (ANSI/ASABE Standards, 2008) for 100 mm thick corrugated vertical cellulose pads. These ow rates were 5, 6.6, 8.3, and
10 L min1 per linear meter of pad mounted vertically. Expressed
in terms of ow rate per exposed surface area of pad (m2 ) in the
test, these numbers were 0.128, 0.171, 0.214, and 0.256 l s1 m2 ,
respectively.
These ow rates were kept constant by varying the pressure of
the hydraulic pump using a power source and adjusting the ow
rate based on readings from the rotameter. The experiment was
also repeated with a dry pad solely to test the pressure drop of the
air passing through it.

222

A. Franco et al. / Computers and Electronics in Agriculture 76 (2011) 218230

Fig. 3. Fluid domains of the tested dry pads: a) 4545 100 mm G&R; b) 4545 100 mm Munters; c) 6030 100 mm Munters; d) 6030 50 mm Munters.

Fig. 4. Thickness of the mean sheet of water of the uid domain of the pad 4545 100 mm G&R when wet (Qw = 0.256 l s1 m2 ).

A. Franco et al. / Computers and Electronics in Agriculture 76 (2011) 218230

223

The application of Darcys law is the standard method for characterizing the homogeneous ow through porous media. For a
permanent ow in a single direction of incompressible and Newtonian uid through a porous horizontal medium, this law is as
follows:

Fig. 5. Diagram of pad meshing 4545 100 mm G&R when dry.

The air ow through the pad was regulated by controlling


the fan speed, taking continuous measurements with a hot-wire
anemometer. The range of airspeed for the test was set between
0.3 and 3 m s1 .
At the start of the test, the water ow was xed. After 10 min,
the fan was started at an initial velocity of approximately 0.3 m s1 ,
increasing by 0.6 m s1 up to 3 m s1 . Increments in speed were
separated by 5 min intervals so that equilibrium could be achieved
between the pad and the new air and water conditions. At each
airspeed, 100 data were recorded by all the sensors at 3 s intervals.
2.3. Aerodynamic analysis of the porous medium
The evaporative pad is a structured porous medium that produces resistance when air passes through it into the greenhouse.
This resistance increases as the pads porosity decreases, usually
due to high water ows moistening the pad or to incrustations of
carbonates and organic remains in the pores when maintenance
is decient. This increase in resistance hampers ventilation and
reduces the efciency of the greenhouse cooling system. It also
implies a higher energetic cost of the fans which make the air
circulate.


P
= V
K
l

(3)

where P is the pressure drop (Pa), l the thickness of the porous


medium (m),  the dynamic viscosity (Pa s), K the Darcys permeability (m2 ) and V the air velocity (m s1 ). However, this law is only
applicable for very low ow rates, since at higher ones the pressure drop is not proportional to the air velocity. The adimensional
number used to characterize the start of non-linear behavior is the
Reynolds number (Re). Non-linear behavior is accepted to occur for
a Reynolds number of between 1 and 10.
At high Reynolds numbers, a strongly inertial regime appears
and Forchheimers empirical equation is used. This equation has
been used by several authors to describe the passage of air through
et al.,
insect-proof screens (Dierickx, 1998; Miguel, 1998a,b; Munoz
1999; Bartzanas et al., 2002; Valera et al., 2006) and is written as:

P

=
V + V 2
Kf
l

(4)

where Kf is Forchheimers permeability (m2 ),  the uid density


(kg m3 ) and the adimensional inertial factor. This equation
denes the pressure drop as the sum of two terms, one of which is
viscous and the other inertial.
Nevertheless, other authors (Amaral and Moyne, 1997; Fourar
et al., 2004) show that the appearance of a weakly inertial regime,
in which the non-linear behavior of ows begins to appear, can be
described by the cubic law:

P
2 3

V
= V+
K

l

(5)

where  is the inertial factor, an adimensional parameter for


weakly inertial regimes. This equation was obtained from numerical simulations in periodical two-dimensional porous media used

Fig. 6. Pressure at the inlet in the models of pads tested dry at an entrance velocity of 0.50 m s1 : a) 4545 100 mm G&R; b) 4545 100 mm Munters; c) 6030 100 mm
Munters; d) 6030 50 mm Munters.

224

A. Franco et al. / Computers and Electronics in Agriculture 76 (2011) 218230

Fig. 7. Retained water volume per surface unit as a function of ow.

Fig. 8. Pressure drop against air velocity in the G&R 4545 100 mm pad both wet and dry (Qw = 0.256 l s1 m2 ).

in homogenization techniques for isotropic and homogenous


media.

2.4. CFD simulation


Computational uid dynamics (CFD) is a sophisticated design
and analysis tool that allows us to study different design limitations
based on a computational model.

The CFD software used in the present work was ANSYS-CFX


from the Workbench package. Its code uses a discretization procedure of nite volumes and it incorporates the equations that
govern heat ow and transfer, such as the continuity equation,
momentum conservation and energy conservation. The overall methodology used to carry out a complete study with the
commercial software used in this work is summarized as follows: generating the geometry of the uid domain, meshing,
dening the boundary conditions and the physical properties of

Table 2
Mean theoretical thickness (mm) of the sheet of water and porosity (m3 m3 ) as a function of the water ow applied.
Pad model

Water ow applied
0 l s1 m2

G&R 4545 100 mm


Munters 4545 100 mm
Munters 6030 100 mm
Munters 6030 50 mm

0.128 l s1 m2
3

e (mm)

 (m m

0.00
0.00
0.00
0.00

0.957
0.959
0.965
0.937

0.171 l s1 m2
3

e (mm)

 (m m

3.59 102
5.82 102
4.93 102
8.25 102

0.943
0.939
0.947
0.891

0.214 l s1 m2
3

e (mm)

 (m m

4.26 102
6.59 102
5.54 102
9.55 102

0.940
0.937
0.945
0.884

0.256 l s1 m2
3

e (mm)

 (m m

4.76 102
7.23 102
5.91 102
10.62 102

0.938
0.934
0.943
0.878

e (mm)

 (m3 m3 )

4.93 102
7.87 102
6.60 102
12.19 102

0.938
0.932
0.941
0.869

A. Franco et al. / Computers and Electronics in Agriculture 76 (2011) 218230

225

Table 3
Coefcients a and b for the best cubic t (P = aV3 + bV), determination coefcient R2 , permeability K and inertial factor () calculated for the four pad models and ve water
ows tested (20).
Qw (l s1 m2 )

R2

K (m2 )

G&R 4545 100 mm

0
0.128
0.171
0.214
0.256

0.820
0.866
0.886
0.966
0.979

5.743
6.388
6.421
6.098
6.316

0.993
0.997
0.997
0.995
0.996

3.221 107
2.896 107
2.881 107
3.034 107
2.929 107

1.089 104
1.151 104
1.177 104
1.283 104
1.301 104

Munters 4545
100 mm

0
0.128
0.171
0.214
0.256

0.586
0.721
0.728
0.786
0.786

4.845
5.222
5.686
5.102
5.679

0.992
0.992
0.996
0.994
0.994

3.818 107
3.543 107
3.254 107
3.626 107
3.258 107

7.786 104
9.580 104
9.673 104
1.044 104
1.044 104

Munters 6030
100 mm

0
0.128
0.171
0.214
0.256

0.759
0.852
0.822
0.854
0.833

6.038
6.153
6.409
6.248
6.629

0.996
0.995
0.996
0.994
0.994

3.064 107
3.007 107
2.887 107
2.961 107
2.791 107

1.008 104
1.132 104
1.092 104
1.135 104
1.107 104

Munters 6030
50 mm

0
0.128
0.171
0.214
0.256

0.625
0.749
0.717
0.791
0.803

4.905
5.685
6.284
5.902
5.879

0.996
0.989
0.988
0.984
0.981

2.810 107
1.562 107
1.413 107
1.505 107
1.510 107

1.230 104
2.073 104
1.985 104
2.189 104
2.223 104

Pad

Fig. 9. Permeability (K) as a function of the porosity of all the pads and water ows tested.

Fig. 10. Inertial factor () as a function of the porosity of all the pads and water ows tested.

226

A. Franco et al. / Computers and Electronics in Agriculture 76 (2011) 218230

Fig. 11. Comparison of simulated and experimental results of pressure drop as a function of the air velocity entering dry pads (Qw = 0 l s1 m2 ): a) G&R 4545 100 mm; b)
Munters 4545 100 mm; c) Munters 6030 100 mm; d) Munters 6030 50 mm.

the model (pre-processing), solving and nally analysis (postprocessing).

2.4.1. Geometry and meshing of the uid domain


The geometry of the uid domain of each pad has been ascertained using computer assisted design (CAD) software by Autodesk.
Using inverse engineering we obtained all the values required
for drawing the geometry of the pad (thickness, curvature radius,
length and width of undulation, etc.) by taking measurements on
different samples of each pad model (Table 1). An axis is drawn corresponding to a sheet, and using equidistances, at the mid-point of
the thickness of the sheet, on both sides of the axis we generate
two new polylines made up of straight lines and circumference
arcs. The two polylines are then joined at the ends to obtain a
closed polygon that will be the basis for generating the solid sheet.
The closed polygon is copied at the suitable distance to form the
other cellulose sheet that will allow us to draw the basic unit of
the model. The two polygons are extruded and we generate solid
entities. The sheets are turned according to the angles of incidence
of the different pads (4545 or 6030 ). Finally, a mould is constructed consisting of a solid parallelepiped with a hollow interior
with the denitive dimensions of each pad. By means of Boolean
operations (fusion of the sheets and difference between the solids
and the mould) we obtain the geometry of the uid domain. To
obtain the air volumes, rst a parallelepiped is generated with the
same dimensions as the hollow interior of the rst mould used
and by measuring the difference between this solid and the one
of the sheets we can obtain the air volume. It is most important
that translational periodicity exists both in the upper and lower
parts of the uid domain, as it will later be applied as a boundary
condition (Fig. 3).
Obtaining the uid domain of the air that passes through the
wet pad is a complex task. Approximating the movement of water
due to gravity over the transfer surface of the pad, we consider that

a mean theoretical sheet of water is a function of the ow applied


and the geometry of the evaporative. A uid model has been made
for the four pads studied with a mean layer of water only for the
maximum ow tested, i.e., 0.256 l s1 m2 . The model of the uid
domain of each type pad when wet was generated from the difference in solids between the two sheets making up the pad, minus
twice the average layer of water running down it (Fig. 4). Once the
geometry of the uid domain was generated, it was imported and
veried using the ANSYS-DesignModeler application.
For the meshing we use the CFX-Mesh application from ANSYS
Workbench, which generates hybrid meshes of tetrahedra and hexahedra. The meshing used was of the non-structured type with
ination on the sides of the model in contact with the sheets of the
cellulose pad, and periodicity at the upper and lower parts of the
solid model, to simulate continuity of the model used at a certain
height (Fig. 5).
2.4.2. Boundary conditions, physical properties and solver
The CFD simulations of evaporative pads carried out in the
present work are steady state, with an isotherm uid domain at
25 C and 1 atm atmospheric pressure. The physical model of turbulence is the Standard k-Epsilon Model. The boundary conditions,
in CFX-Pre, were:
- Inlet: Different speeds were applied to the input area of the model,
with subsonic ow regime and normal speed to said area (V): 0.5,
1, 1.5, 2, 2.5, and 3 m s1 , respectively in each simulation.
- Outlet: Applied to the outlet area of the model, with subsonic ow
regime and average static pressure of 0 Pa.
- Wall: Applied at the sides of the model, in contact with the cellulose sheets, we rene for each limit layer of the mesh, considering
it as a wall which does not slide and smooth wall.
- Domain interface: Applied at the upper and lower areas of the
model, in which periodicity was applied to the geometry and

A. Franco et al. / Computers and Electronics in Agriculture 76 (2011) 218230

meshing of the model, creating domain interfaces, to conserve


the conditions of the uid.
CFX-Solver has been made for a high resolution, with a maximum of 100 iterations, with an automatic control of the scale time
and with a convergence criterion of residual type based on the
residual average of all the control volumes (RMS) at a value of 104 .
2.4.3. Post-processor
Once the CFX-Solver has been applied, we proceed to analyze
the results (post-processor) generating a contour plot that shows
the pressure eld in the inlet area of the model, and obtaining the
average value (Fig. 6). In this way, for each velocity simulated, the
pressure difference between the inlet and outlet is determined for
each type of evaporative pad tested. This difference will now be
compared to the experimental results of the pads obtained in the
wind tunnel.
3. Results and discussion
3.1. Volume of water retained by the pads
The volume of water retained by the evaporative pads per surface unit has been shown to be directly related to the water ow
applied to the pads upper part (Fig. 7). The pad which retained
most water was the Munters 6030 50 mm, between 2.2 and
3.25 l m2 with a maximum error of 3.08%, followed by the Munters
4545 100 mm (2.02 and 2.73 l m2 and a maximum error of
7.74%), the Munters 6030 100 mm pad (1.782.39 l m2 and a
maximum error of 10.03%), and nally the G&R 4545 100 mm
pad (1.401.93 l m2 and a maximum error of 7.63%).
Franco et al. (2010) studied the inuence of water and air ow on
the performance of corrugated cellulose pads, recommending airspeed intervals of between 1 and 1.5 m s1 , obtaining pressure drop
of between 3.9 and 11.25 Pa, depending on the pad thickness and
the water applied. Air saturation efciency was between 64 and 70%
and evaporated water was between 1.8 and 2.62 kg h1 m2 C1 .
Increasing the ow of water applied to the pads, which is directly
related to the water retained in them, did not modify their saturation efciency or the amount of evaporated water, but it did
increase their resistance to the air ow.
3.2. Mean thickness of the sheet of water and porosity of the pads
Based on the results obtained regarding the retained water volume for the 4 types of pads and given the specic area of each
one (Table 1), by applying Eq. (1) we obtain the values of mean
theoretical thickness (e) of the sheet of water which descends the
sheets of the cellulose pads for each ow applied. Knowing these
values we are in a position to generate the geometric model of the
uid domain (air) that passes through the pad for each water ow
applied.
The porosity of the pads () can also be determined as the ratio
between the volume of air (Va ) and the total volume (VT ); or more
easily if we know the solid volume of the cellulose sheets that make
up the pad (Vs ) and the volume of water retained by it (Vl ), previously determined for a known volume of sample (VT ). We can
correlate these in the following equation:
=

Va
VT Vs Vl
Vs
V
=
=1
l
VT
VT
VT
VT

(6)

As may be expected for any geometry, the greater the ow of


water applied to the pad, the greater the volume of water it retains,
the greater the mean theoretical sheet of water descending the pad,
and the smaller the porosity, as Table 2 shows.

227

According to the results obtained, the porosity () is correlated with the geometry of the pads (le /l) and the water ow,
using non-linear regression analysis of the data and obtaining
the following expression, considering that the pad is moistened
(Qw > 0 l s1 m2 ):

 l 0.210

 = 0.215

(1 + Qw 0.029 ),

R2 = 0.881

(7)

3.3. Permeability and inertial factor


In order to determine the permeability and inertial factor, the
four commercial models of evaporative pads have been tested at
ve different water ows. These tests were carried out in a low
velocity wind tunnel which has been described above, obtaining
data pairs that relate each pressure drop produced by the pads at
each water ow with the velocity of the air passing through the
pad. It is known that an increase in air velocity produces a greater
pressure drop (Koca et al., 1991; Liao and Chiu, 2002; Hosseini et al.,
2006). We have observed for all the cellulose pads tested that on
increasing the water ow applied the pressure drop for a given air
velocity also increases. The is due to the reduction in pad porosity
(m3 m3 ) since the increase in water ow increases the sheet of
water descending over the interior transfer surface and reduces
the air volume per unit of volume. This is contrary to the results
obtained by El-Dessouky et al. (1996) for structured pads. Fig. 8
shows the pressure drop for the G&R 4545 100 mm pad against
the air velocity both wet and dry (maximum water ow tested:
0.256 l s1 m2 ).
It can be seen that the two variables (pressure drop an air velocity) are closely related by means of the determination coefcient
(R2 ), using the following form:
P = aV 3 + bV

(8)

Equalling the rst and second terms of the polynomial equation


(Eq. (8)), respectively with the cubic equation (Eq. (5)), the following expressions are obtained to determine permeability (K) and the
inertial factor ():
K=

l
,
b

=

a
l 2

(9)

where l is pad thickness (m), a and b the coefcients of the rst and
second terms of Eq. (8), and taking into account the density  and
dynamic viscosity  of the air for the experimental conditions, the
permeability and inertial factor can be calculated for the four pads
and ve water ows tested (Table 3).
Table 3 shows that the values of permeability tend to increase
as the pad porosity increases, whereas the inertial factor tends to
decrease. Consequently, both parameters can be obtained as a function of porosity. Figs. 9 and 10 show the values of permeability and
inertial factor, respectively, as a function of porosity. The following
equations are the ones that best t the experimental data:
K= 1.53 105  2 + 3.04 105  1.47 105 ,

R2 =0.8153
(10)

=9.53 103  2 18.95 103  9.509 103 ,

R2 =0.9204
(11)

For each pad, the t is not so good, which shows that permeability and inertial factor do not only depend on porosity, but also
on the geometry of the pad. This constitutes the main limitation of
the global approach, and it is here that the approach based on CFD

228

A. Franco et al. / Computers and Electronics in Agriculture 76 (2011) 218230

Fig. 12. Comparison of simulated and experimental results of pressure drop as a function of the air velocity entering wet pads (Qw = 0 l s1 m2 ): a) G&R 4545 100 mm; b)
Munters 4545 100 mm; c) Munters 6030 100 mm; d) Munters 6030 50 mm.

modeling comes into its own, as it takes into account the detailed
geometry of the pad.
3.4. CFD simulation results and verication
The inuence of air velocity on the pressure drop produced
by the different evaporative pads assayed, both wet and dry
(Qw = 0.256 l s1 m2 ) has been obtained experimentally using a
low velocity wind tunnel and simulated by a 3D geometric model
using commercial CFD software (ANSYS-CFX). Comparing the simulated and experimental values of pressure drop at different air
velocities in the four evaporative pads, both wet (Fig. 11) and dry
(Fig. 12), shows high similarity. The percentage difference (%diff)
between the measured and simulated values proposed by Wilson
et al. (2006) is determined by Eq. (12).
%diff =

 MaxP MinP 
MaxP

100%

(12)

We should point out that for dry pads (Qw = 0 l s1 m2 ) these


differences are under 2.19 Pa for air velocities below 2 m s1 , nding a maximum of 3.27 Pa in the Munters 6030 50 mm pad for
an air velocity of 3 m s1 , equivalent to 9.08%diff. Although the differences remain small in wet pads (Qw = 0.256 l s1 m2 ), they are
slightly greater than in dry ones, obtaining differences of under
2.33 Pa for air velocities of less than 2 m s1 , and recording a maximum of 7.10 Pa for the G&R 4545 100 mm pad at a velocity of
3 m s1 , equivalent to 15.53%diff.
The simulated and experimental pressure drop data have been
correlated (Fig. 13) for all the evaporative pads tested both dry and
wet, obtaining very high correlation for dry pads and even higher
correlation for wet ones.
The comparison between experimental data and those simulated by CFD show good correlation when the pads are dry, with
a slope of 1.0143, i.e., an error of 1.43%. This correlation decreased
when the pads were moistened, registering an error of 8.35%. This
error could be due to the fact that the thickness of the sheet of

Fig. 13. Correlation between simulated and experimental pressure drop data for all pads: a) dry (Qw = 0 l s1 m2 ); b) wet (Qw = 0.256 l s1 m2 ).

A. Franco et al. / Computers and Electronics in Agriculture 76 (2011) 218230

water running over the inside surface of the pad was taken to be
constant; although this hypothesis is a good approximation, it may
not be completely true. In the case of higher airspeeds (23 m s1 ),
the differences found were greater. Nevertheless, we consider the
correlation between simulated and measured data to be acceptable,
as in no case did the error exceed 10%.
The CFD simulation procedure carried out in the present work
may prove useful to manufacturers of evaporative pads, who can
improve their design using different geometries in order to enhance
their performance. Indeed, many researchers are currently working on agricultural applications of CFD, a fact witnessed by the
increase in publications on this topic in recent years. We believe
that the methodology applied in this paper may be useful to other
researchers who are interested in enhancing the performance of
similar technologies.
The t between measured and computed pressure drops was
indeed better when using a global modeling based on the porous
medium approach with an experimental determination of permeability and inertial factors. However, this approach requires very
costly experimental equipment.
4. Conclusions
Determining the volume of water retained by the pads is
most useful for determining the porosity and average thickness
of the sheet of water that descends the cellulose pads at different water ows. Good degree of agreement has been found
between porosity and both the water ow applied and the pad
geometry.
Based on the results obtained in the wind tunnel, we can state
that at higher water ow the pressure drop produced by the pad
increases, but to a lesser extent than at higher air ow.The results
obtained for pressure drop against air velocity through the pad at
any water ow applied t a polynomial function known as cubic
equation, with determination coefcients that are very close to the
unit. For porosity of between 0.85 and 0.97 (a range that covers
characteristic values of all the evaporative pads tested at different water ows), the best equation that describes the relationship
between permeability and inertial factor is a second order polynomial.
Comparison of simulated and experimental pressure drop values at different air velocities for the four evaporative pads, both dry
and wet, show a good degree of agreement. The maximum percentage differences are 9.08% for dry pads and 15.53% for wet ones at
air velocity of 3 m s1 .
Modeling of evaporative pads using CFD proves to be a good
tool for optimizing their design, since good correlation was found
between simulated results and those obtained experimentally in
the wind tunnel for both dry and wet pads. Using this technique we can model how an evaporative pad will function for
any air velocity and water ow. We can therefore predict the
pressure drop that the pad will produce and optimize its design
accordingly.
Acknowledgement
The authors wish to express their gratitude to the Junta de
Andaluca (Spain) for partially nancing the present work by means
of the research grants P09-AGR-4593.
References
Al-Helal, I.M., 2007. Effects of ventilation rate on the environment of a fanpad
evaporatively cooled, shaded greenhouse in extreme arid climates. Appl. Eng.
Agric. 23 (2), 221230.
Amaral, H., Moyne, C., 1997. Dispersion in two-dimensional periodic porous media.
Part I. Hydrodynamics. Phys. Fluids 9 (8), 22432252.

229

ANSI/ASABE Standards, 2008. EP406.4: Heating, Ventilating and Cooling Greenhouses. ASABE, St. Joseph, MI.
Baille, A., 2001. Trends in greenhouse technology for improved climate control in
mild winter climates. Acta Hortic. 559, 161168.
Bartzanas, T., Boulard, T., Kittas, C., 2002. Numerical simulation of the airow and
temperature distribution in a tunnel greenhouse equipped with insect-proof
screen in the openings. Comput. Electron. Agric. 34, 207221.
Beshkani, A., Hosseini, R., 2006. Numerical modeling of rigid media evaporative
cooler. Appl. Therm. Eng. 26, 636643.
Baeza, E.J., Perez-Parra, J.J., Lopez, J.C., Montero, J.I., 2006. CFD study of the natural
ventilation performance of a parral type greenhouse with different numbers of
spans and roof vent congurations. Acta Hortic. 719, 33340.
Bartzanas, T., Katsoulas, N., Kittas, C., Sapounas, A.A., 2006. Dispersion of pesticides from a naturally ventilated greenhouse: a CFD approach. Acta Hortic. 718,
07314.
Boulard, T., Haxaire, R., Lamrani, M.A., Roy, J.C., Jaffrin, A., 1999. Characterization
and modelling of the air uxes induced by natural ventilation in a greenhouse.
J. Agric. Eng. Res. 74, 135144.
Boulard, T., Kittas, C., Roy, J.C., Wang, S., 2002. Convective and ventilation transfers
in greenhouses, part 2: determination of the distributed greenhouse climate.
Biosyst. Eng. 83, 129147.
Bournet, P.E., Ould Khaoua, S.A., Boulard, T., 2007. Numerical prediction of the effect
of vent arrangements on the ventilation and energy transfer in a multi-span
glasshouse using a bi-band radiation model. Biosyst. Eng. 98 (2), 224234.
Castilla, N., Hernndez, J., 2005. The plastic greenhouse industry in Spain. Chronica
Hortic. 45 (3), 1520.
Campen, J.B., 2006. Ventilation of small multispan greenhouse in relation to the
window openings calculated with CFD. Acta Hortic. 718, 351356.
Dagtekin, M., Karaka, C., Yildiz, Y., 2009. Performance characteristics of a pad evaporative cooling system in a broiler house in a Mediterranean climate. Biosyst.
Eng. 103, 100104.
Dierickx, I.E., 1998. Flow reduction of synthetic screens obtained with both a water
and airow apparatus. J. Agri. Eng. Res. 71, 6773.
El-Dessouky, H., Al-Haddad, A.A., Al-Juwayhel, F.I., 1996. Thermal and hydraulic
performance of a modied two-stage evaporative cooler. Renew. Energy 7 (2),
165176.
Fang, F.M., Chen, J.C., Hong, Y.T., 2001. Experimental and analytical evaluation of
ow in a square-to-square wind tunnel contraction. J. Wind Eng. Ind. Aerodyn.
89 (34), 247262.
Fatnassi, H., Leyronas, C., Boulard, T., Bardin, M., Nicot, P., 2009. Dependence of greenhouse tunnel ventilation on wind direction and crop height. Biosyst. Eng. 103,
338343.
Fidaros, D., Baxevanou, C., Bartzanas, T., Kittas, C., 2008. Flow characteristics and
temperature patterns in a fan ventilated greenhouse. Acta Hortic. 797, 123130.
Fourar, M., Radilla, G., Lenormand, R., Moyne, C., 2004. On the non-linear behavior of
a laminar single-phase ow through two and three-dimensional porous media.
Adv. Water Resour. 27, 669677.
A., Pena,
A., 2010. Inuence of water and air ow
Franco, A., Valera, D.L., Madueno,
on the performance of cellulose evaporative cooling pads used in Mediterranean
greenhouses. Trans. ASABE 53 (2), 565576.
Hosseini, R., Beshkani, A., Soltani, M., 2006. Performance improvement of gas turbines of Fars (Iran) combined cycle power plant by intake air cooling using a
media evaporative cooler. Energy Convers. Manage. 48., 10551064.
Katsoulas, N., Savas, D., Tsirogiannis, I., Merkouris, O., Kittas, C., 2009. Response of an
eggplant crop grown under Mediterranean summer conditions to greenhouse
fog cooling. Sci. Hortic. 123, 9098.
Kim, K., Yoon, J.Y., Kwon, H.J., Han, J.H., Son, J.E., Nam, S.W., Giacomelli, G.A., Lee,
I.B., 2008. 3-D CFD analysis of relative humidity distribution in greenhouse
with a fog cooling system and refrigerative dehumidiers. Biosyst. Eng. 100,
245255.
Kittas, C., Bartzanas, T., Jaffrin, A., 2003. Temperature gradients in a partially shaded
large greenhouse equipped with evaporative cooling pads. Biosyst. Eng. 85 (1),
8794.
Koca, R.W., Hughes, W.C., Christianson, L.L., 1991. Evaporative cooling pads: test,
procedure and evaluation. Appl. Eng. Agric. 7 (4), 485490.
Lee, I.B., Hong, S.W., Hwang, H.S., Seo, I.H., 2006. Study on ventilation efciencies of naturally ventilated multi-span greenhouses in Korea. Acta Hortic. 719,
341347.
Liao, C.M., Chiu, K.H., 2002. Wind tunnel modeling the system performance of alternative evaporative cooling pads in Taiwan region. Build. Environ. 37, 177187.
A., 2010. Experimental evaluation by
Lopez, A., Valera, D.L., Molina-Aiz, F.D., Pena,
sonic anemometry of airow in a Mediterranean greenhouse equipped with a
padfan cooling system. Trans. ASABE 53 (3), 945957.
Majdoubi, H., Boulard, T., Fatnassi, H., Bouirden, L., 2009. Airow and microclimate
patterns in a one-hectare Canary type greenhouse: an experimental and CFD
assisted study. Agric. Forest Meteorol. 149, 10501062.
Miguel, A.F., 1998. Transport phenomena through porous screens and openings:
from theory to greenhouse practice. Doctoral Thesis. Agricultural University of
Wageningen, Holland. 239 pp.
Miguel, A.F., 1998b. Airow through porous screens: from theory to practical considerations. Energy Build. 28, 6369.
Milosavljevic, N., Heikkil, P., 2001. A comprehensive approach to cooling tower
design. Appl. Therm. Eng. 21, 899915.
Mistriotis, A., Arcidiacono, C., Picudo, P., Bot, G.P.A., Scarascia-Mugnozza, G., 1997.
Computational analysis of ventilation in greenhouses at zero and low wind
speeds. Agric. Forest Meteorol. 88, 121135.

230

A. Franco et al. / Computers and Electronics in Agriculture 76 (2011) 218230

Molina-Aiz, F.D., 2010. Simulacin y modelacin de la ventilacin en invernaderos de


Almera mediante la utilizacin de dinmica computacional de uidos. Doctoral
Thesis. University of Almera, Spain. 868 pp.
Molina-Aiz, F.D., Valera, D.L., lvarez, A.J., 2004. Measurement and simulation of
climate inside Almera-type greenhouses using Computational Fluid Dynamics.
Agric. Forest Meteorol. 125, 3351.
Molina-Aiz, F.D., Fatnassi, H., Boulard, T., Roy, J.C., Valera, D.L., 2010. Comparison of
nite element and nite volume methods for simulation of natural ventilation
in greenhouses. Comput. Electron. Agric. 72, 6986.
Montero, J.I., 2006. Evaporative cooling in greenhouses: effect on microclimate,
water use efciency, and plant response. Acta Hortic. 719, 373384.

Montero, J.I., Munoz,


P., Antn, A., Iglesias, N., 2005. Computational Fluid Dynamic
modelling of night-time energy uxes in unheated greenhouses. Acta Hortic.
691, 403409.

Munoz,
P., Montero, J.I., Antn, A., Giuffrida, F., 1999. Effect of insect-proof screens
and roof openings on greenhouse ventilation. J. Agric. Eng. Res. 73, 171178.
Ould Khaoua, S.A., Bournet, P.E., Migeon, C., Boulard, T., Chasseriaux, G., 2006. Analysis of greenhouse ventilation efciency based on computational uid dynamics.
Biosyst. Eng. 95 (1), 8398.
Roy, J.C., Boulard, T., Lee, I.B., Chave, M., Nieto, C., 2006. CFD prediction of the distribution and deposition of fungal spores in a greenhouse. Acta Hortic. 719,
279286.
Roy, J.C., Vidal, C., Fargues, J., Boulard, T., 2008. CFD based determination of temperature and humidity at leaf surface. Comput. Electron. Agric. 61, 201212.

Sabeh, N.C., Giacomelli, G.A., Kubota, C., 2007. Water use by greenhouse evaporative
cooling systems in a semi-arid climate. In: 2007 ASAE Annual Meeting 074013.
Sethi, V.P., Sharma, S.K., 2007a. Survey of cooling technologies for worldwide agricultural greenhouse applications. Sol. Energy 81, 14471459.
Sethi, V.P., Sharma, S.K., 2007b. Experimental and economic study of a greenhouse thermal control system using aquifer water. Energy Convers. Manage.
48, 306319.
Teitel, M., Ziskind, G., Lirana, O., Dubovsky, V., Letan, R., 2008. Effect of wind direction
on greenhouse ventilation rate, airow patterns and temperature distributions.
Biosyst. Eng. 101 (3), 351369.
Teitel, M., Atias, M., Barak, 2010. Gradients of temperature, humidity and CO2 along
a fan-ventilated greenhouse. Biosyst. Eng. 106, 166174.
Tong, G., Christopher, D.M., Li, B., 2009. Numerical modelling of temperature variations in a Chinese solar greenhouse. Comput. Electron. Agric. 68, 129139.
Valera, D.L., lvarez, A.J., Molina, F.D., 2006. Aerodynamic analysis of several insectproof screens used in greenhouses. Span. J. Agric. Res. 4 (4), 273279.
Valera, D.L., Molina-Aiz, F.D., 2008. Evolucin tecnolgica de los invernaderos. Phytoma 199, 4752.
Wilson, S.D., Dyson, R.W., Tew, R.C., Demko, R., 2006. Experimental and computational analysis of unidirectional ow through stirling engine heater head.
NASA/TM2006-214246, 10 pp.
Willits, D.H., Peet, M.M., 2000. Intermittent application of water to an externally
mounted, greenhouse shade cloth to modify cooling performance. Trans. ASAE
43 (5), 12471252.

You might also like