Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Computers and Electronics in Agriculture 115 (2015) 129–141

Contents lists available at ScienceDirect

Computers and Electronics in Agriculture


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

A method of coupling CFD and energy balance simulations to study


humidity control in unheated greenhouses
D. Piscia, P. Muñoz, C. Panadès, J.I. Montero ⇑
Department of Environmental Horticulture, IRTA, Carretera de Cabrils km 2, 08348 Cabrils, Barcelona, Spain

a r t i c l e i n f o a b s t r a c t

Article history: A coupling method is presented to study the night-time climate in greenhouses. The approach is based on
Received 2 September 2014 two simulation methods: energy balance simulations (ES) and computational fluid dynamics (CFD). The
Received in revised form 7 May 2015 coupled approach takes advantage of the strengths of each method and reduces their weaknesses. Two
Accepted 8 May 2015
CFD parametric studies were carried out. The first one analysed the effects of different wind speeds
and ventilator opening degrees on the ventilation rate. The second study assessed the effects of different
sky temperatures and ventilator opening degrees on convective heat transfer coefficients. The CFD
Keywords:
method was then coupled with the ES through the exchange of two variables: the CFD simulations
Computational fluid dynamics
Energy balance models
provided the ventilation rates for the ES and the CFD method supplied the convective heat transfer coef-
Unheated greenhouses ficients used by for the ES. This coupling approach was applied to the study of night-time ventilation in an
Simulation unheated, polyethylene-covered greenhouse where two important scenarios were addressed: clear-sky
conditions and an overcast sky. Results from the two studies indicate that ventilation during the
night-time in winter improves greenhouse climate; in the clear-sky case, relative humidity was reduced
and the temperature was raised since ventilation reduced or eliminated thermal inversion, whereas in
the covered-sky situation, ventilating reduced the humidity content, but the temperature dropped. As
expected, the increases and drops in temperature and humidity also depended on external conditions.
Minor opening angles produced the greatest changes in terms of greenhouse air temperature and relative
humidity.
Ó 2015 Elsevier B.V. All rights reserved.

1. Introduction conditions. The enormous variety of boundary conditions and


design elements makes greenhouse design a complex task
Greenhouse climate provides a favourable environment for crop (Vanthoor, 2011).
development, so controlling it is crucial for good plant growth. Broadly speaking, greenhouse climate has mainly been simu-
However, this is not a simple matter because several physical lated by means of energy balance simulations (ES) and computa-
phenomena take place (transpiration, condensation, ventilation, tional fluid dynamics (CFD), as described in Kindelan (1980) and
leakage, etc.). In order to improve climate conditions and prevent Boulard et al. (2002), respectively. Zhai and Chen (2005) sum-
the emergence of crop diseases, it is necessary to reduce the effects marised the advantages and drawbacks of these two techniques.
of low temperatures, high humidity and condensation inside the On the one hand, energy balance models, also known in the green-
greenhouse (Bakker, 1991; Baptista et al., 2012). To avoid these house literature as examples of the stirred tank approach (Roy
problems, so-called active (or heated) greenhouses are equipped et al., 2002), are based on the assumption of the uniformity/homo-
with several devices, such as heating, dehumidification and cooling geneity of greenhouse variables such as temperature and humidity.
systems. Meanwhile, in the family of passive (or unheated) green- This assumption makes the implementation of ES computationally
houses, the climate is essentially controlled by means of natural fast and straightforward, but with some major limitations, such as
ventilation (Baeza et al., 2009), which offers little control and the need to have prior information on the empirical expressions for
results in a strong dependence on outside conditions. convective heat fluxes and wind pressure coefficients (Zhang et al.,
Greenhouse performance therefore has to be calculated based on 2013). Furthermore, ES cannot be easily used to describe the effects
all seasonal conditions while considering changes in external of air movements caused by thermal differences or wind. On the
other hand, CFD provides detailed flow patterns and is able to com-
⇑ Corresponding author. Tel.: +34 93 7507511; fax: +34 93 7533954. pute heat transfer fluxes accurately, but is unfortunately much
E-mail address: juanignacio.montero@irta.cat (J.I. Montero). costlier in terms of computational resources, thus making CFD

http://dx.doi.org/10.1016/j.compag.2015.05.005
0168-1699/Ó 2015 Elsevier B.V. All rights reserved.
130 D. Piscia et al. / Computers and Electronics in Agriculture 115 (2015) 129–141

impracticable for exhaustively simulating a fluid dynamics prob- parameters. As reported by Roy et al. (2002), the ventilation rate
lem over a long time period. In addition, fluid equations are stiff of a greenhouse ES model can be computed using Bernoulli equa-
because of the conjugate coupling between wall and air, which tions and values from semi-empirical formulas. The parameters
therefore increases the computational cost of solving the problem. of these semi-empirical equations were derived either from direct
All of these greenhouse climate characteristics make it difficult determination of the discharge coefficients or through in situ deter-
to perform complete greenhouse modelling. Aspects such as mination by doing a regression of an overall coefficient of wind
long-time integration can only be tackled using ES, whereas others, efficiency on ventilation to measure the air exchange rate. In the
such as flow dynamics, can be appropriately described using CFD. present work, the greenhouse under study was an innovative
In this context, coupling CFD with ES could be advantageous: ES structure and no ventilation experiments or specific empirical for-
can make use of the air movement patterns and convective coeffi- mula were known. In order to overcome this problem, a parametric
cients provided by CFD, while ES can provide CFD with the initial study relating the ventilation rate to the air speed and ventilator
values of factors such as the inside temperature and cover opening angle was performed using CFD simulations. Similarly,
temperature. convective heat transfer is governed by a combination of forced
Several studies have addressed the coupling of ES and CFD tech- convection due to wind pressure and free convection caused by
niques in building simulations. Zhang et al. (2013) reported on the buoyancy forces from temperature gradients between the air and
existence of three different types of coupling. The first one is fully different surfaces (walls, cover, soil, plants, etc.). Consequently,
internal coupling, where the set of equations for energy balance convective coefficients depend on the greenhouse type, outside cli-
and CFD are solved together iteratively. In Negrao (1995), it was mate and ventilation conditions. This dependence on several fac-
shown that such internal coupling generates an excessive number tors makes it difficult to choose convective coefficients. For this
of equations for different models (such as ES zones, CFD equations reason, as in the case of ventilation parameters, a parametric study
and plant systems). The second approach is iterative external cou- of the greenhouse was carried out by means of CFD simulations
pling, where the equations of ES and CFD are solved in a segregated and these parameters were then included in the ES.
manner: the variables are exchanged by using an iterative proce- The purpose of this study is twofold. Firstly, it presents and uses
dure until a converged state is achieved. The last type is a method to couple CFD with ES for greenhouse climate modelling:
progressive-replacement external coupling, where the set of vari- CFD simulations were used to obtain ventilation rate formulas and
ables is exchanged after each model reaches a converged state after convective coefficients, which were then included in the energy
each time step. Besides these coupling types, Zhai and Chen (2005) balance mode. Secondly, ES were run to study the effect of ventila-
also described three types of discontinuities between ES and CFD tion on night-time greenhouse climate in terms of humidity, tem-
programmes. The first one is a time-scale discontinuity: ES for heat perature and condensation. The purpose was to address a
transfer in the building enclosure have a characteristic time scale well-known issue: the excess humidity in the night-time in winter
of hours, but CFD uses time steps of seconds for room air. The sec- in unheated greenhouses, given the possibility of thermal inver-
ond is a modelling discontinuity. The indoor values predicted for sions between the interior and exterior temperatures (Montero
each element in ES are spatially averaged, whereas CFD simula- et al., 2004, 2013). Several ES parametric analysis have been per-
tions present field distributions of variables. Finally, the speed dis- formed to assess the effects on the inside temperature and relative
continuity is related to the computational time needed to solve the humidity of different ventilator opening degrees for different out-
model. ES take only a few seconds to perform annual energy anal- side humidity contents.
ysis and require little computer memory, whereas a CFD calcula-
tion for a specific zone may take hours and require considerable
memory. Depending on the type of coupling used, some or all of 2. Materials and methods
these discontinuities may have to be solved.
All the studies on ES-CFD coupling mentioned focus on the In our ventilated greenhouse, two procedures coupled CFD and
energy climate of buildings. Most features of building climate sim- ES: CFD provided the ES with information on ventilation rates and
ulations can also be found in greenhouse climate analysis. In fact, convective heat transfer coefficients. ES was then used to define
the governing equations are the same. However, despite the con- strategies for humidity control based on the management of
siderable similarities between building and greenhouse climate night-time ventilation. Coupling can be summarised in the follow-
modelling, they are different in some ways. One of the most rele- ing steps:
vant differences is occupancy: in a greenhouse, the occupants are
crops; in a building, they are human beings. The inertia loads are 1. A CFD parametric study of the ventilation rate in which only the
also not the same: the typical building envelop presents much momentum equation was considered, so the heat equation and
higher inertia than the greenhouse cover. Another distinction that the mass equation for the water vapour were not included. This
has a major impact on the coupling strategy is the fact that green- study consisted of 24 different combinations of ventilator open-
house climate is strongly linked to exterior conditions because the ing degrees (5°, 10°, 15°, 30°, 60° and 90°) and wind speeds (1,
covering material is semitransparent (plastic or glass) and there- 2, 3 and 4 m s1).
fore participates in radiation. That is why most coupling 2. A CFD parametric study of convective heat transfer in which the
approaches used for buildings are not completely valid for heat equation and the species equation for air and water vapour
greenhouses, such as carrying out separate studies of the indoor were solved. This consisted of 6 CFD simulations for different
and outdoor climates. vent opening angles (0°, 5° and 90°) and sky temperatures
Because of the intimate relationship between indoor and out- (256 and 273 K).
door climate in a greenhouse, the ES-CFD coupling used in this 3. Introduction of the CFD results into the ES in the form of empir-
study was a static bin coupling, as described by Zhai and Chen ical expressions for the ventilation rate and calculated convec-
(2005): the bin coupling process is designed to reduce the comput- tive coefficients.
ing cost. It provides ES the information that is precomputed by CFD
and saved in the bins for continuous energy calculation. CFD sim- After the coupling was performed, ES were used to carry out an
ulated both indoor and outdoor climate and passed information to extensive parametric study of two scenarios (the unheated green-
the ES for correct computation of the ventilation rate and convec- house under clear and overcast skies) because of the low computa-
tive heat transfer because it lacked information on these tional cost of this technique.
D. Piscia et al. / Computers and Electronics in Agriculture 115 (2015) 129–141 131

2.1. The CFD model SIMPLEC method (Van Doormaal and Raithby, 1984), which uses
a relationship between velocity and pressure corrections to enforce
2.1.1. Numerical method mass conservation and obtain the pressure field. Ansys Fluent 13.0
The Numerical Method was the one previously published by the CFD software (Ansys, 2012) was used.
same authors. Below we replicate the explanation given by Piscia
et al. (2013) of such method. 2.1.2. Mesh and boundary conditions
The CFD simulations solved the governing equations of the fluid The CFD model was very similar to the ones in Piscia et al.
flows in the greenhouse. The momentum equation, also known as (2012, 2013), but presented a change in the geometry because of
the Navier–Stokes equation, was derived from the application of the greenhouse ventilator (Fig. 1). The angle of the greenhouse
Newton’s Second Law applied to a fluid element: ventilator was therefore a parameter that changed depending on
! the case. The 3D mesh was unstructured and there were a total
@q U ! ! !
of 443,000 mesh cells. The dimensions of the domain were 160 m
þ U rq U ¼ rp þ lr2 U þSU ; ð1Þ
@x in the x-direction, 50 m in the y-direction, and 12 m in the z-direc-
! tion, chosen in accordance with the criteria of Bournet et al. (2007).
where q is the fluid density, t the time, U the velocity vector, whose
The quality of the grid was checked by applying the skewness aver-
x-component is u, p the pressure, l the turbulent dynamic viscosity
age, which accounts for all the cells. The result was 0.11.
and SU the momentum source.
The CFD used the enhanced wall treatments, which extended
In this situation, the assumption of incompressibility applies
the validity of the near-wall modelling beyond the viscous sub-
and the mass conservation or continuity equation has to be solved:
layer (Ansys, 2012). In the 3D model, the top domain surface was
! defined as a non-slip wall (the sky) and the bottom domain as
qðr U Þ ¼ Sq ; ð2Þ
another non-slip wall that corresponded to the ground. The air flo-
where Sq is the mass source, i.e. crop transpiration in the case of our wed in from the left surface (y–z plane) and flowed out of the
greenhouse. domain through the right. Symmetry was imposed in the domain
The study also involved energy, so the CFD model also had to surfaces (x–y plane) parallel to the wind direction. In addition,
include the heat equation: since our intention was the representation of an infinitely-long
greenhouse, the front walls were not modelled. The 3D mesh
@ qC p T !
þ U rðqcp TÞ ¼ kr2 T þ ST ; ð3Þ was generated using Ansys mesh software.
@t The greenhouse roof and walls were meshed as 0.2-mm thick
where T is the temperature, k the thermal conductivity, ST the heat elements. These surfaces were considered to be semi-transparent
source and cp the heat capacity at constant pressure. solids and their optical properties for far infrared radiation were
Turbulence was modelled using the standard k–e model as follows: absorptivity 0.69, transmissivity 0.19, and reflectivity
(Launder and Spalding, 1972). This method is based on the solution 0.12. These values were assumed to be independent of the wave-
of two equations: one for k, which accounts for turbulent kinetic length. Furthermore, the soil was modelled as a grey media with
energy, and the other for et, where the rate of dissipation of energy emissivity 0.98 for black mulching, following Liakatas et al.
in unit volume and time is considered. This is probably the most (1986). The heat transfer from the soil surface to the greenhouse
used and validated turbulence model, and in the greenhouse CFD was the soil heat flux (SHF), which was set as a boundary
literature it has been used in many research studies (Boulard and condition.
Wang, 2000).
The contribution of radiation was added as a source component 2.1.3. CFD model used for the ventilation rate parametric study
(ST) in Eq. (3). The Discrete Ordinate Model (DOM) was used to cal- The CFD ventilation rate parametric study simulated 24 differ-
culate the radiation component because plastic is a participating ent situations in which only the momentum equation, Eq. (1),
medium and DOM is recommended for semitransparent materials was activated. Since the objective of this parametric study was to
(Verstaag and Malalasekera, 1995; Baxevanou et al., 2008). Details compute the equilibrium ventilation rate, the simulations were
of the DOM can be found in Piscia (2012) and Piscia et al. (2012). run with constant parameters to obtain a final steady state (see
In order to link the fluid density with the other variables, an Table 1). Note that the ventilation studied was windward.
additional equation was necessary, which in our case was the ideal
gas law. In terms of species transport equations and conservation 2.1.4. CFD model used for the convective heat transfer parametric
equations for air and water vapour, the local mass fraction is pre- study
dicted through the solution of a convection–diffusion equation for In addition to the momentum equation, this set of CFD simula-
each species (Ansys, 2012). tions also took into account the energy equation, Eq. (3), as well as
The lettuce (Lactuca sativa) crop inside the greenhouse was con- the water vapour balance equation, which includes crop transpira-
sidered to be homogeneous and a constant water vapour source at tion and condensation. This parametric study was based on 6 CFD
night with a production rate of 1.74  106 kg m2 s1 (Piscia et al., simulations, where the CFD model simulated two extreme ventila-
2012). Condensation was also accounted for in the CFD simulations tor opening degrees (5° and 90°). The aim was to obtain the con-
by means of a user-defined function (UDF). The condensation vective coefficients for these two cases and to use a linear
model applied in this study was developed by Bell (2003). This interpolation to extract the coefficients between these two points.
model was implemented as UDF into the general CFD simulation For each situation, two different sky temperatures were simulated
code. In Bell’s model the condensation rate is governed by the rate to test the effect of different sky values on convective heat coeffi-
of diffusion of water vapour towards the cold surface. The conden- cient values. Finally, the same approach was applied to a closed
sation rate was included in Eqs. (2) and (3) as part of the corre- greenhouse for the sake of comparison because, when a green-
sponding source term in each case. Further details can be found house is closed, the coefficients are driven by buoyancy, whereas
in Piscia (2012) and Piscia et al. (2012). when the greenhouse is open, convection can be modified by the
The convective terms in the CFD equations were modelled using wind. As shown in Table 2, only one wind speed, uW = 2 m s1,
a second-order up-wind scheme, while the viscous term used a and one SHF were considered because of the computational cost
second-order central scheme and the transient term used an impli- limitation. The convective coefficients and their relations were
cit scheme. Pressure–velocity coupling was resolved by the only used in the ES under these conditions.
132 D. Piscia et al. / Computers and Electronics in Agriculture 115 (2015) 129–141

Fig. 1. Three-dimensional view of greenhouse geometry with parametric roof ventilator opening.

Table 1 The energy balance model used in this study is described in


Boundary conditions for the ventilation rate parametric study. Piscia et al. (2013). The ES model established the energy balance
of the cover, side walls, soil surface as well as the energy and mass
Boundary conditions Type
balance of the greenhouse air. No separate energy balance was
Top domain surface Non-slip wall
made for the windward and leeward sidewalls. The thermal inertia
Bottom domain surface Non-slip wall
Left domain surface Inlet velocity 1, 2, 3, 4 m s1 at a height of 2 m of greenhouse elements was not considered. However, the equa-
(logarithmic profile) (Richards and Hoxey, 1993) tions proposed there did not account for ventilation.
Right domain surface Outlet boundary conditions Consequently, the relevant equations A.18 and A.19 in Appendix
Parallel domain surface Symmetry A in Piscia et al. (2013) had to be modified. The energy balance
Ventilator opening degrees 5°, 10°, 15°, 30°, 60°, 90°
of greenhouse air had to include the ventilation term:

V qðcp þ wair cp;x ÞdT air


¼ As qc;s þ Asxi qc;sxi þ Aci qci
dt
Table 2 þ Uðcp T air þ cp;w wair T air þ he wair
Boundary conditions for the convective heat transfer parametric study.
 cp T e  cp;w we T e  he we Þ; ð4Þ
Boundary conditions Type
Top domain surface Sky temperature 256 K, 273 K where the subscripts air and e refer to the greenhouse air and exte-
Bottom domain surface Greenhouse soil heat flux 25 W m2 rior air, w represents the humidity ratio, cpw the heat capacity at
Left domain surface Inlet temperature 276 K and velocity constant pressure for water vapour, U the ventilation rate, V the
2 m s1 at a height of 2 m (logarithmic
volume of air inside the greenhouse and he the enthalpy of conden-
profile) (Richards and Hoxey, 1993)
Right domain surface Outlet boundary conditions sation. Additionally, subscripts c, s, swi, and ci stand for convective,
Parallel domain surface Symmetry soil, sidewall interior and cover interior, while A and q represent
Ventilator opening degrees 0°, 5°, 90° areas and heat fluxes. The convective coefficient listed in Table 3
Initial conditions Values ° were used in the calculation of the heat fluxes by convection q.
Exterior air temperature 276 K
Relative humidity of exterior air 80%
Likewise, the mass balance of greenhouse air can be expressed as:
Inside air temperature 280 K
Relative humidity of inside air 68% dM W
¼ Ccrop  Xcov þ Uðwair  we Þ ð5Þ
Side wall temperature 280 K dt
Cover temperature 280 K
where Mw is the mass of water vapour, C the transpiration rate and
X the condensation rate on the greenhouse cover.

2.2. Energy balance simulations


3. Results
ES are based on the resolution of energy and mass conservation
in a large volume. In the greenhouse literature, energy balance 3.1. CFD ventilation rate parametric study
models are also referred to as perfectly stirred tank models. This
approach requires uniform temperature, humidity and CO2 content The CFD greenhouse model described in Section 2.1.3 studied
inside the greenhouse and uses a ‘‘big leaf’’ model to treat the plant six different vent opening degree configurations (5°, 10°, 15°, 30°,
canopy and describe the exchanges of latent and sensible heat with 60°, 90°) and four different wind speeds (1, 2, 3, 4 m s1), resulting
the inside air (Roy et al., 2002). in a total of 24 scenarios. Ventilation rate was deduced by

Table 3
Convective coefficients for different sky temperatures and roof ventilator opening angles.

a (W m2 K1) (Tsky, K)


(273,90) (256,90) (273,5) (256,5) (273, 0) (256,0)
Interior cover 5.2 5.2 3.4 3.5 2.2 2.3
Exterior cover 3.9 3.9 3.8 3.7 9.4 9.3
Interior sidewall 8.9 8.7 5.1 5.1 1.6 1.8
Exterior sidewall 4.0 4.1 6.8 6.7 6.0 5.9
Soil in greenhouse 12.4 12.2 7.4 7.0 3.1 3.1
D. Piscia et al. / Computers and Electronics in Agriculture 115 (2015) 129–141 133

integrating the velocity field over the vent openings. The relevant the correlation coefficient was r2 = 0.98. The coefficients were
results are shown in Fig. 2. b = (4.0 ± 8.7) 103 kg s1 m1 and c = (1361.0 ± 3.2) 103 kg m2,
The ventilation rate seemed to increase linearly with the sine of whose errors were much smaller than the former ones. With Eq.
the ventilator opening angle for the four wind speeds. The same (9), the maximum relative error was e  22% and the average rela-
linearity reported by Jong (1990) between the ventilation rate, tive error was e  10%, which were lower than the previous
the sine of the angle and the velocity was observed in our equations.
greenhouse. In all these methods, the error for low ventilation rates was
The simplest mathematical way to express the data is as quite high. In order to achieve greater accuracy, a quadratic spline
follows: was computed to capture the behaviour of the ventilation rate as a
function of wind speed and the sine of the vent opening angle.
U ¼ a uw þ b sin K; ð6Þ
Details of the spline are given in the Appendix A. The spline pro-
2 2
where a = (30.3 ± 8.4) 10 kg m and b = (23.7 ± 3.9) 101 vided the best fit for the data: maximum relative error was
kg m1 s1 were obtained using a least-squares method fit. e  11% and average relative error was e  3%. As a consequence,
However, there was a major error between this fit and the original this calculation reduced the error. It is also clear that when com-
data: a maximum relative error of e  240% and an average relative pared with the other methods the spline produced the best esti-
error e  67%. A potential reason was that the simplest equation mate. The spline equation was used by the ES model to compute
was not the most general one to express the linear relationship the ventilation rate. This represented the coupling between CFD
between the ventilation rate, wind velocity and the sine of the ven- and ES, which, according to the coupling types defined in Zhang
tilator opening. The most general equation is, in fact, the following: et al. (2013), is a form of external, static coupling.
U ¼ a uw þ b sin K þ cuw sin K; ð7Þ
3.2. Parametric study of CFD convective coefficients
Furthermore, for the sake of consistency with our CFD simula-
tions, the ventilation rate had to be zero if the greenhouse was The CFD greenhouse model, described in Section 2.1.4, studied
closed (K = 0°), regardless of wind velocity. Therefore, a had to be six different cases, a combination of two different equivalent sky
equal to zero and the data had to be fit in this equation: temperatures (273 K and 256 K) and three different ventilator
opening degrees (0°, 5° and 90°). These temperatures were chosen
U ¼ b sin K þ cuw sin K; ð8Þ
to study the phenomena occurring in unheated greenhouses (ther-
Note that when uw or sin K was constant, the ventilation rate mal inversion and humidity) during cold and clear-sky nights
was linear to the other magnitude. Additionally, Eq. (8) introduced (256 K) and covered skies at an intermediate temperature
a crossed term that could be useful to account for mild nonlinear- (273 K). For the sake of comparison, the two extreme cases of ven-
ities. These coefficients were computed using the least-squares tilation (small and high vent opening angles) and the closed green-
method once again, thus yielding b = (0.4 ± 12.5) 102 kg m1 s2 house were analysed. The convective coefficients of the six
and c = (136.1 ± 4.6) 102 kg m2. In fact, using Eq. (8) provided a simulations are listed in Table 3.
better fit for the data, thus indicating the need to include the The following conclusions can be drawn from Table 3 with
crossed term cuw sin K. In this scenario, the maximum relative respect to the different convective coefficients.
error was e  29% and the average relative error was e  13%.
The only problem with this fit was the large error in coefficient 3.2.1. Convective coefficient for the internal surface of the cover
b, probably induced by the fact that for a two-variable fit, 24 points The coefficient value was the highest for the fully open ventila-
is a rather small set of data. tor (90°) and the lowest for the closed greenhouse. This trend
Another way to estimate these coefficients but with smaller agrees with the fact that convection in the inner cover must
errors was as follows. When the wind velocity was constant become more relevant when the roof is opened. For each ventilator
(Fig. 2), Eq. (8) could be rewritten: opening angle, two sky temperatures were used, but the results
show that the inner convective coefficients had little correlation
U ¼ ðb þ cuw Þ sin K  m sin K; ð9Þ
with this parameter.
As a consequence, by performing linear regressions that went In addition, it is important to compare these values with the
through the origin on the plots of Fig. 2, four m values were ones presented in the greenhouse climate literature. Since the bib-
obtained. An additional regression between m and uw had to be liography contains no descriptions of convective coefficients for
performed to determine the other coefficients. In the worst case, greenhouses with an open ventilator, the comparison was only
possible with the convective coefficients for closed greenhouses.
For comparison purposes, the inner cover CFD coefficients were
6
expressed as a = aDT0.33, where a is a constant and DT is the
5 1 m/s difference in temperature between the cover and the air in the
2 m/s greenhouse. For instance, for a = 2.3 W m2 K1, CFD simulations
4 provided DT = 1.8 K, thus resulting in a = 1.9DT0.33. This expression
3 m/s
(kg s-1)

is between the one given by De Halleux (1989), a = 1.86DT0.33, for a


3 4 m/s
large-scale greenhouse, and the one provided by Papadakis et al.
(1992), a = 2.21DT0.33, for a small polyethylene-covered green-
2
house. Additionally, these convective coefficients could be
1 estimated via the Nusselt number, Nu, because a = kair Nu/L, where
L is the characteristic length of the process and
0 kair = 0.025 W m1 K1 the thermal conductivity of air. In Roy
0 0.2 0.4 0.6 0.8 1
et al. (2002), it is suggested using the roof slope length for the
sin greenhouse cover, so we used L = 3.39 m. The Nusselt number
Fig. 2. Ventilation rate as a function of the sine of the vent opening angle for
depends on the type of convection. For a closed greenhouse, con-
different values of wind speed. Straight lines are the result of fitting the data with vection is driven solely by density gradients because of the absence
U = m sin K. external forcing. This convection mode is called free convection
134 D. Piscia et al. / Computers and Electronics in Agriculture 115 (2015) 129–141

and, depending on the Grashof number, Gr, the flow is considered greenhouse cover, so we used L = 3.39 m. As in Section 3.2.1, the
laminar (Gr < 108) or turbulent (Gr > 108): Nusselt number was useful to estimate greenhouse convective
coefficients. For the external surface of the cover, forced convection
gbDTL3 was the driving mechanism because of the presence of wind,
Gr ¼ ; ð10Þ
v2 whose velocity was considered as the characteristic velocity of
where b = 3.1  103 K1 and v = 1.4  105 m2 s1 are the volumetric the process. The Reynolds number was Re = 4.8  105, so the flow
thermal expansion coefficient and kinematic viscosity at 10 °C, was clearly turbulent and Eq. (13) could be used. Therefore, the
respectively. Considering DT = 1.8 K and typical values of the Nusselt number was Nu = 1136.6, yielding a = 8.4 W m2 K1,
parameters for air, the Grashof number resulted in Gr = 1.2  1010, which was also in good agreement.
thus being turbulent. In this scenario, the Nusselt number was com- When the greenhouse was opened, the convection mode for the
puted via the following expression: external surface of the cover had to be the same. Of course, U and L
must be different, but the formulas from De Halleux (1989) and
1=3
Nu ¼ 0:14ðGrPr Þ ; ð11Þ Papadakis et al. (1992) should be valid. Let us assume L = 2.39 m,
because the roof slope length is halved due to opening the ventila-
where Pr represented the Prandtl number. Considering Pr = 0.714,
tor. Then, these formulas provided a reasonable range for the char-
yielded Nu = 288.7, which in turn provided a good estimate for
acteristic velocity, U = 0.17–0.71 m s1. Taking into account these
the convective coefficient, a = 2.1 W m2 K1.
velocities, the Reynolds number range was Re = 2.9  104, to
When the greenhouse is open, convection became strongly
Re = 1.2  105 corresponding to a laminar and a turbulent flow,
influenced by wind velocity, thus becoming a forced convection.
respectively. The following expression provides the Nusselt num-
In this case, the Reynolds number, Re, determines whether the flow
ber in case the forced-convection flow is laminar:
is laminar (Re < 5  104) or turbulent (Re > 5  104):
UL Nu ¼ 0:67 Re1=2 Pr1=3 ð14Þ
Re ¼ ; ð12Þ
v 1
For U = 0.17 m s , the flow was supposed to be laminar, yield-
where U represents the characteristic velocity inside the green- ing Nu = 102.5, so a = 1.1 W m2 K1, which was a poor estimate of
house. Wang et al. (1999) estimated the interior air speed in a ven- the values from Table 3, a = 3.7–3.9 W m2 K1. Nevertheless, con-
tilated greenhouse as the ratio between the mean ventilation flux sidering U = 0.71 m s1 (turbulent flow), resulted in Nu = 374.6 and
and the section perpendicular to the direction of this flux. a = 3.9 W m2 K1. This value was in good agreement with Table 3
Following their reasoning and using the data from Section 3.1, it and supported the assumption that the flow was turbulent.
was possible to estimate U. For instance, for 90° we obtained
U = 1.1 m s1, so Re = 2.7  105 and the flow from the forced convec- 3.2.3. Convective coefficient for the internal surface of the sidewall
tion was turbulent. Note that when the greenhouse is open, the roof The lowest convective coefficient for the inner sidewall was the
slope length increases for the internal surface of the cover. For 90°, one for the closed greenhouse. As expected, when the greenhouse
this length resulted in 4.39 m, which was considered as L. Following was closed, the air movement occurring inside the greenhouse by
Roy et al. (2002), the Nusselt number is: convection was less than in the case of the ventilated greenhouse.
It was no surprise that when the ventilator was open, the coeffi-
Nu ¼ 0:036 Re4=5 Pr 1=3 ð13Þ
cient had to increase as the greenhouse became more ventilated.
which in our case resulted in Nu = 866.4, thus yielding As in the previous coefficients, the sky temperature had little effect
a = 4.9 W m2 K1, in good agreement with the value on the convective coefficients.
a = 5.2 W m2 K1 from Table 3. This case is similar to that of the coefficient of the internal sur-
face of the cover. Therefore, following the same reasoning,
3.2.2. Convective coefficient for the external surface of the cover a = 1.35DT0.33. In Tantau (1975), a = 1.247DT0.33 is reported for
In this case, the highest convective coefficient corresponded to the convective coefficient of a vertical wall and was quite close
the closed greenhouse, which makes sense. In fact, it was more to the value obtained in the CFD simulations.
than twice the amount in the ventilated situations. Once the ven- For the internal surface of the sidewall, the span height was
tilator was opened, the difference between 5° and 90° was very considered as the characteristic length, L = 3.5 m. The resulting
slight. The differences between the closed and ventilated green- Grashof number was Gr = 1.8  1010, so it was turbulent (Gr > 108).
house arose due to the fact that, by opening the roof ventilators, Following the same procedure as in Section 3.2.1, the Nusselt num-
the greenhouse geometry changed and, as a result, the outside ber was Nu = 328.1 and a = 2.3 W m2 K1, which was close to
air pattern also changed. The presence of ventilators made the a = 1.8 W m2 K1 reported in Table 3.
overall air speed decrease, along with the convective coefficient By means of opening the greenhouse, the former free convec-
of the external surface of the cover. This behaviour is difficult to tion became forced. Following the same reasoning as for the inter-
take into account using the semi-empirical formulas found in the nal surface of the cover, the same estimate for the characteristic
greenhouse research literature. As with the inner cover coeffi- velocity was used, thus U = 1.1 m s1 for 90°. Considering the
cients, sky temperature had an insignificant effect on the convec- same characteristic length, L = 3.5 m, a Reynolds number for a tur-
tive coefficients. bulent flow was obtained, Re = 2.75  105. The resulting Nusselt
The convective coefficients of the CFD outer cover could be number was Nu = 722.8, so the convective coefficient was
compared to the values reported in other papers. For the closed a = 5.2 W m2 K1. This value was quite close to the values for 5°
greenhouse, the CFD convective coefficient was approximately and different but of the same order as the ones for 90°. Note that
9.4 W m2 K1 for a wind speed of 2 m s1. This value was between the same characteristic length was considered for a closed and
8.27 W m2 K1, which was given by De Halleux (1989) for a open greenhouse, which is not necessarily true.
large-scale greenhouse, and 10.4 W m2 K1, which was provided
by Papadakis et al. (1992) for a polyethylene-covered greenhouse 3.2.4. Convective coefficient for the external surface of the sidewall
with uw 6 6.3 m s1. In order to use the expression proposed in As far as we have observed, this convective coefficient has the
De Halleux (1989), a characteristic length L had to be employed. most complex dynamics. The lowest convective coefficient was
Roy et al. (2002) suggests using the roof slope length for the found for the fully open ventilator (90°), whereas the highest
D. Piscia et al. / Computers and Electronics in Agriculture 115 (2015) 129–141 135

was for 5°. When the greenhouse was closed, the coefficient was When the greenhouse was open, the convection mode changed
slightly lower than the value for 5°. This was curious in that exte- from free to forced. As in Sections 3.2.1 and 3.2.3, the same charac-
rior convection on the sidewalls seemed to be slightly enhanced at teristic velocity inside the greenhouse for 90° was considered,
low ventilation angles. As with the former coefficients, the sky U = 1.1 m s1, so, Re = 7.9  103–7.9  104 depending on L. The flow
temperature had little relationship with the convective coeffi- was laminar for L = 0.1 m, so Eq. (14) yields Nu = 53.1, while con-
cients. The CFD convective coefficients for the outer cover were sidering L = 1 m implied the utilisation of Eq. (13) because the flow
comparable to the values reported in other papers. was turbulent, Nu = 265.3. In the end, we obtained
For the closed greenhouse, the CFD convective coefficient was a = 13.3 W m2 K1 in the laminar case, and a = 6.63 W m2 K1
approximately 6 W m2 K1 for a wind speed of 2 m s1. This value in the turbulent flow. This was a clear indication that the flow close
was of the same order as the value for a large-scale greenhouse to the soil was laminar.
given by De Halleux (1989), 8.1 W m2 K1.
As it was done for the external surface of the cover, convection 3.2.6. Coupling between CFD and ES
was forced by the wind and the convective coefficient was esti- As discussed throughout this section, the convective coefficients
mated using the Nusselt number. Taking L = 3.5 m, as the charac- changed due to the ventilation opening angle. The values shown in
teristic length, the Reynolds number was Re = 5  105, thus being Table 3 could be included in the ES model, but there was no esti-
a turbulent flow. In this case, Nu = 1166, so the convective coeffi- mation for the convective coefficients when the ventilator opening
cient was a = 8.3 W m2 K1, which is different from the values degrees were between 5° and 90°. Performing additional CFD sim-
reported in Table 3 but in the same level of agreement as the ulations for the whole range of angles seemed appropriate but too
semi-empirical formula in the former paragraph. costly in terms of computational resources. In addition, there was
Following the same reasoning as in Section 3.2.2, when the the possibility that the results were not that sensitive with the con-
greenhouse was open, the convection mode for the external sur- vective coefficients because of the strong influence of the external
face of the sidewall had to be the same. The characteristic length meteorological conditions. This means that in principle for high
was considered the same as before, while the U was supposed to ventilation rates, the climate inside the greenhouse would be the
be different because the flow was severely altered respect to the same as the external one, no matter which convective coefficients
closed case. Nevertheless, the formula given by De Halleux were considered. Meanwhile, for small ventilator opening degrees,
(1989) should be valid and could be used to provide an estimate the dynamics inside the greenhouse should not be entirely driven
of the characteristic velocity, U = 0.83 m s1 for the fully open ven- by the external conditions and the values of the convective coeffi-
tilator (a = 4 W m2 K1). The resulting Reynolds number was cients might still be relevant.
Re = 2.1  105, corresponding to a turbulent flow. This value of Re Therefore, we considered K = 5°, and studied how changing the
yielded Nu = 577.5 and a = 4.1 W m2 K1, which were consistent values of the convective coefficients influenced the temperature
with our assumptions. and relative humidity of the air inside the greenhouse for the
two scenarios: cold (Tsky = 256 K) and intermediate (Tsky = 273 K).
The remaining parameters are compatible with the ones used for
3.2.5. Convective coefficient between inside air and soil surface
the CFD simulations (Table 2). The initial temperatures of inside
This coefficient followed the same trend as the convective coef-
air, sidewall, soil and cover were the same,
ficients for the internal surface of the sidewall: indeed it was pos-
Tair = Tsw = Ts = Tc = 280 K, while the exterior temperature was
itively correlated to the quantity of air entering the greenhouse.
Te = 276 K. The internal and external values of relative humidity
The convective coefficient for the fully ventilated greenhouse was
were RHair = 68% and RHe = 80%, respectively. Wind velocity was
four times greater than the closed scenario. Once again, the sky
2 m s1 and the soil heat flux was 25 W m2 K.
temperature did not substantially affect these convection
In each case, an energy balance simulation was run for K = 5°
coefficients.
and the corresponding values of the convective coefficients from
Considering that DT = 2.9 K, where DT was the difference
Table 3, and the results of these simulations were stored.
between the soil and inside temperature, the soil coefficient could
Afterwards, a convective coefficient was changed to an intermedi-
be expressed as a = 2.18DT0.33. This value was between the for-
ate value between 5° and 90°. For instance, for the internal part of
mula a = 1.86DT0.33 given by De Halleux (1989) for a large-scale
the cover, a changes from 3.5 to 4.35 W m2 K1. This procedure
greenhouse and a = 3.4DT0.33 reported in Stoffers (1985) for a
was repeated for each a and the results were compared with the
screened greenhouse. It was thus a reasonable agreement.
default value (Table 4). It is clear that interior climate conditions
As with the previous coefficients, a could be estimated by
did not change substantially, except for the relative humidity in
means of the Nusselt number, but an estimate for L is necessary.
the cold case, where it increased about DRHair = 1.8%. However, this
Following Roy et al. (2002), it is suggested that the characteristic
error was acceptable according to Baptista (2007) because it was of
length is of a few centimetres for a bare soil and a few metres
the same order associated with ES. Consequently, the exact values
for a soil covered with plastic film. In our case, the soil was covered
of the convective coefficients were not that relevant, and an esti-
with a black mulching, which is in between both scenarios, so we
mate of them could provide reasonable results without the neces-
took L = 0.1–1 m. Free convection occurred inside the greenhouse
sity of running extensive CFD simulations. Therefore, a linear
when it was closed. Using Eq. (10), a range of Grashof numbers
was obtained, Gr = 5.1105–5.1  108, corresponding to a laminar
and a turbulent flow, respectively. The following expression pro- Table 4
vides the Nusselt number in case the free-convection flow is Sensitivity of the Tair and RHair under changes in the convective coefficients for K = 5°.
laminar: a (W m2 K1) cold/ Cold (Tsky = 256 K) Intermediate
intermediate (Tsky = 273 K)
1=4
Nu ¼ 0:54 ðGr PrÞ ð15Þ
DTair (K) DRHair (%) DTair (K) DRHair (%)
On the one hand, in case the flow was laminar (L = 0.1 m), aci: 3.5/4.35 2.0  101 +1.8 4.6  102 +2.8  101
Nu = 13.3, so a = 3.3 W m2 K1, which was a good estimate of ace: 3.7/3.8 +1.7  102 +1.8  101 +5.5  105 6.9  104
the values from Table 3, a = 3.1 W m2 K1. On the other hand, aswi: 5.1/6.9 3.8  102 +1.5  101 +5.5  105 6.9  104
aswe: 6.7/5.4 6.3  102 +1.6  102 +2.2  103 1.4  102
considering the flow turbulent, Nu = 99.8 and a = 2.5 W m2 K1,
as: 7.0/9.6 +2.7  102 1.8  101 +9.9  102 6.0  101
which was also in good agreement.
136 D. Piscia et al. / Computers and Electronics in Agriculture 115 (2015) 129–141

interpolation as a function of vent opening was implemented temperature is lower than the outside temperature and the relative
between 5° and 90°. Of course, when the ventilators were closed, humidity is higher than the recommended maximum. The effects
the coefficients were taken as the CFD values obtained for 0°. of different ventilator openings on the inside climate depended
This action represented the second coupling link between the on outside climate conditions, which is why the parametric study
CFD and ES models because it evidenced that it was not necessary simulated a combination of different ventilator opening degrees
to run further CFD simulations and provided good estimates of the and different external humidity ratios. A set of 225 ES was run with
convective coefficients for the ES. different external conditions (0.001 6 w 6 0.004) and ventilator
openings (0° 6 K 6 90°). The results were plotted as
3.3. ES parametric study three-dimensional graphics in order to assess the effects of
ventilation for different external humidity ratios. For the sake of
In a greenhouse, it is important to protect crops from low tem- comparison, the closed greenhouse scenario was included in these
peratures and high humidity. In unheated (passive) greenhouses, plots (K = 0°).
the climate is essentially controlled by ventilation. Our aim was Considering an exterior temperature of 276 K, it is clear from
to propose ventilation strategies based on ES parametric analysis Fig. 3 that opening the ventilators prevented thermal inversions
in two different scenarios: from occurring. The greatest effects in terms of temperature gain
were obtained in the first opening degrees. After 10–15°, the
 Section 3.3.1: A cold case, where the inside temperature was effects on the inside temperature were greatly reduced and, after
colder than the outside (unheated greenhouse and clear-sky approximately 30°, the derivative of the temperature in terms of
conditions). the opening was close to zero. In fact, as the opening angle
 Section 3.3.2: An intermediate case, where the inside tempera- increased, thermal inversion was less likely to occur because the
ture was warmer than the outside, but the relative humidity internal and external temperatures tended to balance each other
was high (unheated greenhouse and overcast-sky conditions). out. As expected, the inside temperature did not vary as a result
of the external humidity ratio.
The first scenario was chosen because a previous study (Piscia Likewise, Fig. 4 shows that the relative humidity inside the
et al., 2012) showed that thermal inversion occurred in an greenhouse was always reduced when the ventilator was opened.
unheated greenhouse (SHF 25 W m2) under clear-sky conditions Even an external relative humidity of 85% (corresponding to out-
(256 K). This is because the greenhouse cover emits more infrared side values of 276 K and w = 0.004) was less than the situation in
radiation than it receives from the sky. During a clear winter night, closed greenhouses, which have a relative humidity of almost
the equivalent sky temperature can be 20 °C lower than the air 90%. As occurred for the temperature (Fig. 3), most of the effects
temperature and the cover can consequently be up to 3 °C cooler in terms of RH reduction were obtained within the first 10–15°
than the outside air. The second scenario was proposed because of ventilator opening. If the outside air was very humid, ventilation
Piscia et al. (2012) reported that in this situation thermal inversion reduced the greenhouse RH very little. Indeed, opening the green-
is prevented but the humidity can be higher than the maximum house to reduce the relative humidity only made sense when the
tolerable threshold (85%, as suggested by Campen, 2009). Of outside humidity was lower than the inside. It can be inferred from
course, both are major problems for crop production. To provide the ES parametric study that it is always useful to ventilate in the
realistic results, the ES model used the ventilation rate and convec- event of thermal inversion. This conclusion agrees with results
tive coefficients found in the CFD simulations presented in Sections reported by Baptista (2007).
3.1 and 3.2. The condensation rate in terms of the outside air humidity and
the ventilator opening angle is shown in Fig. 5. It can be concluded
3.3.1. Cold case: equivalent sky temperature of 256 K and soil heat flux that when the external humidity was high, ventilation did not pre-
of 25 W m2 vent condensation. In some cases, condensation actually drastically
As explained above, this scenario can occur on cold, clear-sky increased. The presence of high external humidity and a high ven-
nights in unheated greenhouses. In this situation, the inside tilation rate combined with a low cover temperature paved the

Fig. 3. Inside temperature for different combinations of ventilator opening degrees and external humidity ratios for the case of an equivalent sky temperature of 256 K and a
soil heat flux 25 W m2.
D. Piscia et al. / Computers and Electronics in Agriculture 115 (2015) 129–141 137

Fig. 4. Inside relative humidity for different combinations of ventilator opening degrees and external humidity ratios for the case of an equivalent sky temperature of 256 K
and a soil heat flux of 25 W m2.

Fig. 5. Condensation rate for different combinations of ventilator opening degrees and external humidity ratios for the case of an equivalent sky temperature of 256 K and a
soil heat flux 25 W m2.

way for a high condensation rate. The condensation rate can be that fall in this category are more difficult to handle because any
overpredicted by the ES model because this numerical technique possible gain of a lower humidity content could be lost by letting
provides a single result. Meanwhile, in the CFD simulations, only warmer air escape from the greenhouse, thus lowering the temper-
the water vapour in contact with the surface below the dew point ature of the cover and the interior air. As in Section 3.3.1, a para-
temperature was able to condensate; in the ES model, all water metric study was performed in the covered-sky scenario in terms
vapour below the dew point temperature condensated. Although of ventilation and the exterior humidity ratio. A total of 225 ES
the condensation rate might be overestimated in quantity com- cases was simulated with different exterior humidity ratios
pared to the real scenario, its results are clear from a qualitative (0.001 6 w 6 0.004) and openings ((0° 6 K 6 90°).
point of view. Ventilation reduced the inside humidity but did In the intermediate case there was no thermal inversion.
not necessarily prevent condensation from forming on the inner Therefore, in terms of heat conservation, opening the ventilator
roof. always implied a drop in temperature, regardless of the humidity
ratio of the outside air, as shown in Fig. 6. In addition, air exchange
3.3.2. Intermediate case: equivalent sky temperature of 273 K and soil with only a 5° opening angle had a relevant impact on overall
heat flux of 25 W m2 greenhouse temperature. For higher angles of ventilator opening,
The case of the covered sky presented major differences com- the interior temperature tended asymptotically towards the out-
pared to clear-sky conditions. Under a covered sky, thermal inver- side temperature of 276 K.
sion does not occur because the equivalent sky temperature is We compared the relative humidity in a covered-sky scenario
almost the same as the outside air temperature. This is related to (Fig. 7) with the clear-sky scenario (Fig. 4). In all cases, the relative
the high humidity level that greenhouse air can reach. Situations humidity dropped when the vents were opened slightly. Of course,
138 D. Piscia et al. / Computers and Electronics in Agriculture 115 (2015) 129–141

Fig. 6. Inside temperature for different combinations of ventilator opening degrees and external humidity ratios for the case of an equivalent sky temperature of 273 K and a
soil heat flux of 25 W m2.

Fig. 7. Inside RH for different combinations of ventilator opening degrees and external humidity ratios for the case of an equivalent sky temperature of 273 K and a soil heat
flux of 25 W m2.

Fig. 8. Condensation rate for different combinations of ventilator opening degrees and external humidity ratios for the case of an equivalent sky temperature of 273 K and a
soil heat flux of 25 W m2.
D. Piscia et al. / Computers and Electronics in Agriculture 115 (2015) 129–141 139

due to the gradient of water vapour concentration between the experimental measurements (Piscia et al., 2012, 2013). Three tech-
inside and the outside, the relative humidity tended to balance niques are commonly used (the heat balance, mass balance or tra-
out, so it dropped when the outside RH was lower. In this scenario, cer gas technique and the decay method using nitrous oxide) and a
the main drawback of opening vents was that the inside tempera- full review of them can be found in Roy et al. (2002).
ture was reduced, as cooler air flowed into the greenhouse. In con- The ES model can be seen as a surrogate form of the CFD model
clusion, to prevent excessive humidity problems in the crop, (Eldred et al., 2004) that can quickly include CFD predictions in a
ventilating the greenhouse is recommended in all cases except much smaller time scale. In this perspective, the ES represents a
when the outside humidity is over 85%, i.e. the threshold value way of making the CFD results available outside the CFD frame-
given by Campen (2009). A good choice is an opening angle of work. The ES model can easily be used by a wider audience that
between 10° and 15° because the RH can be reduced enough in includes growers and technical advisors.
most cases without losing too much heat. In addition, the CFD model did not take into account the ques-
Regarding condensation rate for covered skies, it can be tion of air leakage. In CFD modelling, taking infiltration into
observed from Fig. 8 that condensation was always prevented by account is difficult, since it requires defining the location and
opening the ventilator. This is because the humidity content of geometry of the openings through which the greenhouse air could
0.004 kg kg1 is below the value obtained by computing the exit and this information is not usually available. However, venti-
humidity ratio at the saturation point imposed by the cover tem- lation can be compared with known cases of air leakage. For
perature. Consequently, there can be no condensation in any case. instance, if an air leakage value of 1.34 V h1 (López et al., 2001;
Baille et al., 2006) is considered for a wind speed of 2 m s1, then
4. Discussion the air leakage corresponds to a ventilation rate of 0.04 kg s1,
which is one order of magnitude smaller than the ventilation rate
This paper introduces a novel way of using CFD and ES models provided by a 5° ventilator opening (0.3 kg s1). Even though air
to study greenhouse climate. The idea was not to take an exclusive leakage can be neglected when the greenhouse is ventilated, this
approach, but to allow the two models to collaborate and gain a second-order effect can be included in the ES models by adapting
better understanding of the system. In this perspective, collabora- the semi-empirical formulas found in the literature (López et al.,
tion was based on the exchange of CFD convective coefficients and 2001; Baille et al., 2006).
ventilation rates with the ES model. The coupled ES model had the In the future, the ES model can be enhanced by coupling it with a
advantage of being independent of empirical coefficients, thus plant response model, which would make it possible to directly link
making it more robust. the effects of different ventilation strategies to plant production out-
The coupling approach presented here can be improved by per- puts. This further step is in the same direction as the one proposed by
forming only one comprehensive CFD parametric study to extract Vanthoor (2011), which presented a global design model that took
the convective coefficients and ventilation rates, instead of two dif- into account energy, plant response and economic returns.
ferent sets of simulations. The problem with this approach is that it As stated above, we only studied the case of an unheated green-
has higher computational costs, but the advantage is that it pro- house because our aim was to study a Mediterranean-based green-
vides a ventilation rate that takes into account the air movement house. Nevertheless, Piscia et al. (2012) reported that humidity
caused by the combined effect of wind and buoyancy. This study was not reduced through ventilation of a heated greenhouse; in
presented a subset of all possible climate scenarios. The convective fact, the inside temperature was higher than the outside tempera-
coefficients are particularly applicable for unheated conditions and ture and the relative humidity was below the threshold limit of
one wind speed and direction. Nevertheless, the approach was 85%. These results may seem surprising, since it is a common prac-
modular and further studies should be done to include heated sit- tice in Central European greenhouses to combine heating and ven-
uations as well as all or part of the spectrum of wind direction and tilation or to force preheated external air for humidity control
wind speed. Including more parameters and scenarios will greatly (Campen, 2009). This may be due to the fact that the value for
increase the number of CFD simulations required. In order to night-time transpiration was taken from a previous study in the
reduce the number of simulations, design of experiment tech- unheated greenhouse and was considered a constant. Night-time
niques can be a used. The aim is to create an experimental design transpiration is probably higher in heated greenhouses.
that can sample a high-dimensional space in a representative way Therefore, results from the ES and CFD models may differ from
with a minimum number of samples. Classical examples of design the aforementioned data from Piscia et al. (2012). In any case,
of experiments include central composite design (Giunta et al., modelling night-time transpiration in heated greenhouses is a
2003), which requires 1 + 2N + 2N samples, and the Box–Behnken research subject beyond the scope of the present study.
design, which calls for 1 + 4N + (N  1)/2 samples. However, these Based on the analysis, for the scenarios taken into account,
CFD simulations were not necessary in our case because the cou- night-time ventilation is recommended in both clear and overcast
pled method proposed in this paper was used to develop ventila- sky conditions. This conclusion agrees with the findings of Baptista
tion strategies to improve night-time climate conditions in (2007). Moreover, another advantage of night-time ventilation is
passive greenhouses in terms of humidity, temperature and con- that it can significantly reduce the onset of Botrytis cinerea disease
densation for a wide range of situations. (Baptista et al., 2012). Besides ventilation, there are other ways of
The coupling presented in this paper was a one way transfer of improving greenhouse climate conditions. For instance, as reported
information from CFD to ES; it is possible to have a two-way in Piscia et al. (2013), it would be advantageous for the cover to be
exchange of information between the models by selecting the most made of a material with high reflectivity (low emissivity and trans-
relevant cases identified by the ES parametric study and running missivity). Unfortunately, no such material is currently available.
detailed CFD simulations of those selected cases. This was not Another reasonable option would be the use of materials to accu-
undertaken in the present study in order to limit the extension mulate heat during the day and release this energy at night, thus
of this paper. increasing the greenhouse temperature. For example, containers
Validation is a fundamental part of a simulation model, but was filled with water are an economic way of storing heat. However,
not included in this paper due to the intrinsic difficulty of trying to more expensive procedures may be available such as using
measure the greenhouse ventilation rate accurately. The CFD sim- phase-change materials to increase the greenhouse temperature.
ulations, ES models and their coupling had been previously vali- As with water containers, the soil could also be used for heat stor-
dated in our experimental setting with good agreement with age. Throughout this paper, the soil is considered a surface with
140 D. Piscia et al. / Computers and Electronics in Agriculture 115 (2015) 129–141

constant conductivity, diffusivity and water content. Changing soil B4 ¼ maxð0; uw  2Þ; ðA5Þ
properties could be a useful way to improve the soil’s thermal stor-
age capacity, which could be beneficial for greenhouse climate. B5 ¼ maxð0; 2  uw Þ; ðA6Þ

5. Conclusions B6 ¼ B4 maxð0; sinðKÞ  0:499770102643102Þ; ðA7Þ

This paper introduces a new approach to greenhouse climate


B7 ¼ B4 maxð0; 0:499770102643102  sinðKÞÞ; ðA8Þ
simulation. The method is based on the coupling of an ES model
B8 ¼ maxð0; 0:865759839492344  sinðKÞÞ; ðA9Þ
with a CFD model. This technique was applied to study the effect
of ventilation on greenhouse temperature, humidity and
B9 ¼ B8 maxð0; 2  uw Þ; ðA10Þ
condensation.
Two CFD parametric studies were performed to compute the
ventilation rates and convective coefficients. The first study Appendix B. Nomenclature and subscripts
obtained the ventilation rate for different combinations of wind
speeds and ventilator opening angles. The results showed strong
linearity between the ventilation rate and the ventilator opening A area (m2)
angle and were included in the ES model by means of a quadratic a fitting parameter (kg m2)
spline. The second CFD study computed the convective coefficients b fitting parameter (kg s1 m1)
for different external conditions. The CFD simulations show that c fitting parameter (kg m2)
convective coefficients depend strongly on the ventilation rate, cp heat capacity at constant pressure of air (J kg1 K1)
but do not suffer substantial variations when the sky temperature cp,w heat capacity at constant pressure of water vapour
is changed. (J kg1 K1)
This work analysed the effects of ventilation on greenhouse cli- CFD computational fluid dynamics
mate under different conditions, specifically two scenarios of DOM discrete ordinate model
major interest. The first represented clear-sky conditions and the ES energy balance simulation
second overcast-sky conditions combined with a lack of heat he enthalpy of condensation of water (J kg1)
(SHF of 25 W m2). Under clear-sky conditions, the coupled ES k turbulent kinetic energy (J)
model indicated that ventilation was always advisable, but that L characteristic length (m)
opening the roof ventilators slightly was sufficient. Under Mw mass of water vapour (kg)
overcast-sky conditions, ventilation was recommended in most m slope of the linear fit
cases. Ventilation was only unnecessary when the outside air N number of variables
humidity was over 85%. Further studies could combine ventilation p pressure (Pa)
with the use of containers of water or different soils with higher q heat flux (W m2)
conductivity as ways of increasing the efficiency of passive r linear correlation coefficient
greenhouses. RH relative humidity (%)
SU momentum source (kg m2 s2)
Acknowledgements ST heat source (W m3)
Sq mass source (kg m3 s1)
The work presented here was carried out within INIA (project SHF soil heat flux (W m2)
RTA2012-00039-CO2-01). This research was also supported by t time (s)
the EUPHOROS Project, the European Commission, the T absolute temperature (K)
Directorate General for Research, the 7th Framework Programme UDF user-defined function
for RTD, Theme 2 e Biotechnology, Agriculture & Food, contract ! velocity vector (m s1)
211457. Thanks are also given to INIA fellowship FPI-INIA 92. U
u velocity component of the x coordinate (m s1)
uw wind velocity (m s1)
Appendix A. Quadratic spline
V greenhouse volume (m3)
w humidity ratio (kg kg1)
The ventilation rate values obtained by the CFD parametric
a convective heat transfer coefficient (W m2 K1)
study were embedded in the ES model by using the following
quadratic spline:
e relative error (%)
e average relative error (%)
U ¼ ð19:3941877273305 þ 15:0399995903575 B1 et rate of dissipation of the turbulent kinetic energy
 22:5664934534249 B2  4:50069762381828 B3 (m2 s3)
k thermal conductivity (W m1 K1)
þ 9:85686151491185 B4  14:6458277948187 B5 l turbulent dynamic viscosity (kg m1 s1)
þ 19:1931416709916 B6  19:418993497707 B7 q fluid density (kg m3)
 15:074984018779 B8 þ 16:4915503604416 B9Þ=12; ðA1Þ K ventilator opening angle (°)
DT temperature difference (K)
where U ventilation rate (kg s1)
C transpiration rate of the crop (kg s1)
B1 ¼ maxð0; sinðKÞ  0:258690844053802Þ; ðA2Þ
X condensation rate on the cover (kg s1)
Subscript
B2 ¼ maxð0; 0:258690844053802  sinðKÞÞ; ðA3Þ
air greenhouse air
B3 ¼ B1 maxð0; uw  2Þ; ðA4Þ c convective
D. Piscia et al. / Computers and Electronics in Agriculture 115 (2015) 129–141 141

ci cover interior Kindelan, M., 1980. Dynamic modeling of greenhouse environment. T. ASAE 23 (5),
1232–1239.
e exterior air Launder, B.E., Spalding, D.B. (Eds.), 1972. Lectures in Mathematical Models of
s soil Turbulence. Academic Press, London.
swi sidewall interior Liakatas, A., Clark, J., Monteith, J., 1986. Measurements of the heat balance under
plastic mulches. Part I: radiation balance and soil heat flux. Agric. Forest
sky sky Meteorol. 36 (3), 227–239.
López, J.C., Pérez, J., Montero, J.I., Antón, A., 2001. Air infiltration rate of Almeria
parral type greenhouses. Acta Hortic. 559 (1), 229–232.
Montero, J., Munoz, P., Anton, A., Iglesias, N., 2004. Computational fluid dynamics
modelling of night-time energy fluxes in unheated greenhouses. International
References Conference on Sustainable Greenhouse systems (special issue). Acta Hortic. vol.
691, pp. 403–410
Ansys, 2012. User guide 13.0. Lebanon, NH, USA. Montero, J.I., Muñoz, P., Sánchez-Guerrero, M.C., Medrano, E., Piscia, D., 2013.
Baeza, E.J., Pèrez-Parra, J.J., Montero, J.I., Bailey, B.J., Lòpez, J.C., Gàzquez, J.C., 2009. Shading screens for the improvement of the night time climate of unheated
Analysis of the role of sidewall vents on buoyancy-driven natural ventilation in greenhouse. Span. J. Agric. Res. 11 (1), 32–46.
parral-type greenhouses with or without insect screens using computational Negrao, C.O.R., 1995. Conflation of Computational Fluid Dynamics and Building
fluid dynamics. Biosyst. Eng. 104 (1), 86–96. Thermal Simulation. PhD Thesis. University of Strathclyde.
Baille, A., Lopez, J.C., Bonachela, S., Gonzalez-Real, M.M., Montero, J.I., 2006. Night Papadakis, G., Frangoudakis, A., Kyritsis, S., 1992. Mixed, forced and free
energy balance in a heated low-cost plastic greenhouse. Agric. Forest Meteorol. convection heat transfer at the greenhouse cover. J. Agric. Eng. Res. 51, 191–
137 (1), 107–118. 205.
Bakker, J.C., 1991. Analysis of Humidity Effects on Growth and Production of Piscia, D., 2012. Analysis of Night-Time Climate in Plastic-Covered Greenhouses.
Glasshouse Fruit Vegetables. PhD Thesis. Agricultural University of PhD Thesis. Universitat Politècnica de Catalunya.
Wageningen. Piscia, D., Montero, J.I., Baeza, E., Bailey, B.J., 2012. A CFD greenhouse night-time
Baptista, F.J.F., 2007. Modelling the Climate in Unheated Tomato Greenhouses and condensation model. Biosyst. Eng. 111 (2), 141–154.
Predicting Botrytis cinerea Infection. PhD Thesis. Universidade de Evora. Piscia, D., Montero, J.I., Bailey, B., Muñoz, P., Oliva, A., 2013. A new optimisation
Baptista, F.J., Bailey, B.J., Meneses, J.F., 2012. Effect of nocturnal ventilation on the methodology used to study the effect of cover properties on night-time
occurrence of Botrytis cinerea in Mediterranean unheated tomato greenhouses. greenhouse climate. Biosyst. Eng. 116, 130–143.
Crop Prot. 32, 144–149. Richards, P.J., Hoxey, R.P., 1993. Appropriate boundary conditions for computational
Baxevanou, C., Bartzanas, T., Fidaros, D., Kittas, C., 2008. Solar radiation distribution wind engineering models using the k-e turbulence model. J. Wind Eng. Ind.
in a tunnel greenhouse. Acta Hortic. 801 (1), 855–862. Aerod. 46, 145–153.
Bell, B., 2003. UDF for Condensation of Steam from Moist Air (Available from Ansys Roy, J.C., Boulard, T., Kittas, C., Wang, S., 2002. PA—precision agriculture: convective
Fluent UDF repository). Ansys Fluent Inc.. and ventilation transfers in greenhouses, part 1: the greenhouse considered as a
Boulard, T., Wang, S., 2000. Greenhouse crop transpiration simulation from external perfectly stirred tank. Biosyst. Eng. 83 (1), 1–20.
climate conditions. Agric. Forest Meteorol. 100 (1), 25–34. Stoffers, J.A., 1985. Energy fluxes in screened greenhouses. Instituut voor
Boulard, T., Kittas, C., Roy, J.C., Wang, S., 2002. SE-structure and environment: Mechanisatie. Arbeid en Gebouwen. Wageningen (Lecture Ag Eng 84 NIAE
convective and ventilation transfers in greenhouses, Part 2: determination of Silsoe).
the distributed greenhouse climate. Biosyst. Eng. 83 (2), 129–147. Tantau, H.J., 1975. Der Einfluss von einfach und Doppelbedachungen auf das Klima
Bournet, P., Ould Khaoua, S., Boulard, T., 2007. Numerical prediction of the effect of und den Wärmehaushalt von Gewächshäusern (Effects of single and double
vent arrangements on the ventilation and energy transfer in a multi-span covering on climate and heat management in greenhouses). PhD Thesis. Fak.
glasshouse using a bi-band radiation model. Biosyst. Eng. 98 (2), 224–234. Gartenbau und Landeskultur Tech. Univ. Hannover.
Campen, J.B., 2009. Dehumidification of Greenhouses. PhD Thesis. Agricultural Van Doormaal, J.P., Raithby, G.D., 1984. Enhancements of the SIMPLE method
University of Wageningen. for predicting incompressible fluid flows. Numer. Heat Transfer 7 (2), 147–
De Halleux, D., 1989. Dynamic Model of Heat and Mass Transfer in Greenhouses: 163.
Theoretical and Experimental Study. PhD Thesis. Gembloux. Vanthoor, B., 2011. A Model-Based Greenhouse Design Method. PhD Thesis.
Eldred, M.S., Giunta, A.A., Collis, S.S., 2004. Second-order corrections for surrogate- Agricultural University of Wageningen.
based optimization with model hierarchies. In: Proceedings of the 10th AIAA/ Verstaag, H., Malalasekera, W. (Eds.), 1995. An Introduction to Computational Fluid
ISSMO Multidisciplinary Analysis and Optimization Conference. August 30– Dynamics: The Finite Volume Method. Longman Group.
September 1, 2004. Albany (New York) USA. AIAA Paper 2004–4457. Wang, S., Boulard, T., Haxaire, R., 1999. Air speed profiles in a naturally ventilated
Giunta, A.A., Wojtkiewicz Jr., S.F., Eldred, M.S., 2003. Overview of modern design of greenhouse with a tomato crop. Agric. Foest Metorolog. 96 (4), 181–188.
experiments methods for computational simulations. In: Proceedings of the Zhai, Z.J., Chen, Q.Y., 2005. Performance of coupled building energy and CFD
41st Aerospace Sciences Meeting and Exhibit. January 6–9, 2003. Reno simulations. Energy Build. 37 (4), 333–344.
(Nevada), USA. AIAA Paper 2003–649. Zhang, R., Lam, K.P., Yao, S.C., Zhang, Y., 2013. Coupled EnergyPlus and
Jong, T., 1990. Natural Ventilation of Large Multi-Span Greenhouses. PhD Thesis. computational fluid dynamics simulation for natural ventilation. Build.
Agricultural University of Wageningen. Environ. 68, 100–113.

You might also like