Benni 2016

download Benni 2016

of 46

Transcript of Benni 2016

  • 8/16/2019 Benni 2016

    1/46

    Accepted Manuscript

    Title: Ef ficacy of greenhouse natural ventilation:environmental monitoring and CFD simulations of a study

    case

    Author: Stefano Benni Patrizia Tassinari Filippo Bonora

    Alberto Barbaresi Daniele Torreggiani

    PII: S0378-7788(16)30375-9

    DOI:   http://dx.doi.org/doi:10.1016/j.enbuild.2016.05.014

    Reference: ENB 6649

    To appear in:   ENB

    Received date: 2-10-2015

    Revised date: 13-4-2016

    Accepted date: 6-5-2016

    Please cite this article as: Stefano Benni, Patrizia Tassinari, Filippo Bonora,

    Alberto Barbaresi, Daniele Torreggiani, Ef ficacy of greenhouse natural ventilation:

    environmental monitoring and CFD simulations of a study case, Energy and Buildings

    http://dx.doi.org/10.1016/j.enbuild.2016.05.014

    This is a PDF  file of an unedited manuscript that has been accepted for publication.As a service to our customers we are providing this early version of the manuscript.

    The manuscript will undergo copyediting, typesetting, and review of the resulting proof 

    before it is published in its  final form. Please note that during the production process

    errors may be discovered which could affect the content, and all legal disclaimers that

    apply to the journal pertain.

    http://dx.doi.org/doi:10.1016/j.enbuild.2016.05.014http://dx.doi.org/10.1016/j.enbuild.2016.05.014http://dx.doi.org/10.1016/j.enbuild.2016.05.014http://dx.doi.org/doi:10.1016/j.enbuild.2016.05.014

  • 8/16/2019 Benni 2016

    2/46

    1

    Efficacy of greenhouse natural ventilation:

    environmental monitoring and CFD simulations of a study case

    Stefano Benni*, Patrizia Tassinari a , Filippo Bonora b ,

     Alberto Barbaresi c , Daniele Torreggiani d

    Department of Agricultural Sciences, University of Bologna. Viale Fanin 48, 40127Bologna (Italy)

    *Corresponding author, email: [email protected]

    phone +39 051 2096166 fax +39 051 2096171

    a) [email protected]

    b) [email protected]

    c) [email protected]

    d) [email protected]

    mailto:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]

  • 8/16/2019 Benni 2016

    3/46

    2

    Graphical Abstract

    Highlights

      CFD analyses of natural ventilation in an horticultural greenhouse were

    performed

      Cooling achievable through various configurations of vents opening of

    was assessed

      Closed windward roof vent and open wall vent entailed 64% of maximum

    heat removal  The other possible scenarios showed a performance index of about 50%

      Results suggest to enhance vent control system considering wind

    direction as input

  • 8/16/2019 Benni 2016

    4/46

    3

    Abstract

    Indoor microclimate control is fundamental in greenhouse design, and vent

    dimensions and positions play a crucial role in natural ventilation management.This research considers an Italian greenhouse for horticultural production and

    aims at identifying optimal vent configurations and opening management

    procedures for indoor environment control, focusing on summer cooling.

    Numerical modelling of airflows and temperature distributions was carried out

    through finite element CFD software, with streamline upwind discretization

    schemes for advection terms. Calibration of the numerical modelling was

    performed by comparing data collected in controlled environmental condition

    with simulations results. The automatic vent opening system of the greenhouse

    is programmed to fully open all the windows of each span when indoor air

    temperature overcomes a threshold value. Numerical simulations were

    performed to assess the efficacy of this solution in comparison with alternative

    strategies. Various configurations of roof vents were tested, with side wall vents

    always open. The best performances were obtained with windward roof vent

    closed, which entailed 64% of the maximum heat removal achievable through

    natural ventilation. The other possible scenarios considered showed a

    performance index of about 50%. The results suggest therefore to enhance the

    vent control system by considering also wind direction as input.

    Keywords: CFD simulation; Airflow; Greenhouse; Cooling

  • 8/16/2019 Benni 2016

    5/46

    4

    1.  Introduction

    The distribution of environmental parameters in greenhouses should be

    maintained as much uniform as possible for general purposes of crop production

    [1]. In particular, most of the plants grown in greenhouses fit at an optimal

    temperature between 17°C and 27°C, with extremes of 10°C and 35°C [2]. Up to

    outdoor temperatures of 27°C, internal temperature control can be achieved by

    means of ventilation, while for higher external temperatures (30-35°C for

    extended periods) it is necessary to intervene by means of artificial cooling in

    order to maintain microclimate functional to the physiology of cultivated plants.

    Moreover, inlet of outdoor air is essential to replenish the CO2  consumed by

    plants for the photosynthesis process. Indoor microclimate control is thus a

    fundamental issue in greenhouse design and analysis [3 –6], and natural

    ventilation plays a crucial role in indoor climate control [7], as it directly affects

    heat and mass transport between the outside environment and the greenhouse,

    thus strongly influencing its inside climate [8]. Vent dimensions and positions are

    key elements in natural ventilation design. The correlations between ventilation

    rates and environmental parameters, including wind speed and direction, havebeen assessed with various approaches [9]. A thermal model based on energy

    balance was proposed by [10] to predict the natural ventilation rate of

    greenhouses. Decay-rate tracer of N2O techniques were used by Kittas et al. [11]

    to experimentally investigate the influence of vent type and screening on airflow

    and temperature distribution within a mono-span greenhouse with round-arch

    roof and vertical walls. It was found that in case of side openings only, indoor air

    velocity showed a fast flow near the ground and low velocity near the roof, while

    the combination of roof and side openings caused an increase in air velocity and

    a decrease in indoor temperature, together with a greater microclimate

    heterogeneity. Demrati et al. [12] computed global energy balance of a large-

    scale greenhouse in hot climate, based on the values of inside and outside air

    temperature and humidity, outside global solar radiation, and wind speed and

    direction measured in significant time spans, thus determining ventilation fluxes.

    Airspeed measurements through vents and inside the greenhouse were also

  • 8/16/2019 Benni 2016

    6/46

    5

    performed to compute airflow patterns. In this regard, Kittas et al. [13] collected

    data through hot-wire anemometry to deepen previous analyses on natural

    ventilation in relation to vent opening choices. In particular they focused on

    greenhouses ventilated by both roof and side vents and provided experimental

    results for the main greenhouse climate parameters (including air velocity and air

    temperature) in the whole greenhouse volume, resulting in a database for

    validating computational fluid dynamics (CFD) research approaches for

    greenhouse ventilation. A discussion on the efficiency of different discretization

    methods used as CFD solvers for the simulation of natural ventilation in

    greenhouses was developed by Molina-Aiz et al. [14]. Ayuga [15] underlined that

    ventilation is one of the major problems in greenhouse technology, and a good

    balance between airflow and air speed can be designed by means of proper CFD

    analyses. Teitel et al. [16] underlined the importance of analysing air movements

    in greenhouses, caused by ventilation, and their effects on indoor microclimate

    uniformity. In fact growers has been increasingly tending to exploit every area of

    the greenhouse for high quality yields, due to the ever stronger competition

    within the global market. Experiments - developed by assessing the decay rate of

    a tracer gas (N2O), and by means of 3D sonic anemometer and dry and wet-bulb

    thermocouples - showed that ventilation rate increases linearly with wind speed,

    and gradients in temperature and air velocity occurred in the vertical and

    horizontal directions. CFD simulations were extensively used to investigate the

    effects of structure shape, and ventilator size and arrangement (with or without

    insect-proof screens) on microclimate [17]. Teitel and Wenger [18] compared

    experiments, CFD, and a model relating flow rate through openings to their

    pressure drops, to determine the air exchange in a small naturally ventilated

    monospan greenhouse. The three methods agreed well up to a wind speed of

    about 4 m/s, while at higher wind speeds the ventilation rates were lower, with

    consequently diminished air-exchange efficiency, due to imperfect mixing of the

    supply air with the greenhouse air. The complexity and realism of both

    experimental analyses and simulations has been continuously increased, as

    recent studies account for all the relevant variables concerning the greenhousesystem [19]. The effects of vent configurations in terms of natural ventilation

  • 8/16/2019 Benni 2016

    7/46

    6

    have been considered as a topical subject in several researches. Bartzanas et al.

    [8] numerically investigated the effects of ventilation configuration of a tunnel

    greenhouse on airflow and temperature patterns was through CFD with

    validation against experimental data. Results indicated that the highest

    ventilation rates are not always the best criterion to assess the performance of

    different ventilation systems in greenhouses. Analyses must consider also the air

    velocities in the zone interested by crops, the efficiencies of ventilation on flow

    rate, and the air temperature differences between inside and outside.

    Furthermore [20] showed that the basic mechanism of heat transfer in a

    greenhouse is the convection due to the entering air stream, also when the

    global solar radiation of a sunny day is considered. Sun radiation represents the

    greatest heat load for a greenhouse in summer periods in Europe. For example in

    the area of Bologna, in Centre-Northern Italy there is an average of 2074 annual

    hours of sunshine and 250 h of monthly sunshine from May to September [21]. 

    Nevertheless, in certain conditions of hour angle, airflows and temperature

    distributions in a greenhouse can be analysed independently on the

    consideration of the external incident radiation [22]. This study deals with the

    thermo-fluid-dynamic analysis of a greenhouse in these conditions.

    Specifically, this research considers a study case of a common greenhouse for

    horticultural production and aims at identifying optimal vent configurations and

    vent opening management procedures for indoor environment control, focusing

    on summer cooling. The specific objective is the assessment of the efficacy of

    natural ventilation strategies in greenhouse cooling, in environmental conditions

    which allow target temperatures to be reached by means of natural ventilation,

    with no HVAC system. In particular, the study aims at assessing the effect of

    natural ventilation in the conditions when the contribution of solar radiation can

    be neglected.

  • 8/16/2019 Benni 2016

    8/46

    7

    2.  Materials and methods

    2.1.  Greenhouse description

    The experimental greenhouse is a 307 m2  three-span greenhouse of the

    Department of Agricultural Sciences of the University of Bologna, located in the

    Imola branch of the University (44.337340° N latitude, 11.718647° E longitude

    and 72 m altitude, about 30 km far from Bologna). The azimuth of spans

    longitudinal axis is 55° direction, with access doors located on the north-eastern

    front of the building. Each span is 8 m wide, 12.7 m long, and 4 m high at eave,

    with 5.5 m maximum height (pitch slope is 40%). The greenhouse has concrete

    floor and steel structure, with 4 mm thick coverings of tempered glass. Each span

    has two central top roof openings of 25° over 1.50 m length. Two wall openings

    of the same size are located at the top of the north-western and south-eastern

    side walls (Figure 1). Both roof and side vents are continuous along the whole

    length.

    The spans correspond to indoor areas separated by glass walls and connected

    through internal doors. This study focuses on the south-eastern span (on the left

    in Figure 1). This sector is provided with heated benches with aluminium

    structure, lamps for supplementary photoperiodic lighting, a reverse osmosis

    system for the treatment of irrigation water, and a high-precision mixer for the

    formulation of fertilizer solutions. The areas is dedicated to experimental and

    educational activities related to propagation, protection, nutrition, and

    programming of phenological phases of potted and soilless ornamental plants.

    Moreover this sector is used to experience directly microclimate management in

    a protected environment and test the interactions between climate parameters

    and plants growth. The equipment allows indoor environment of each sector to

    be managed independently, by means of a computer which controls the heating

    and cooling systems, vent opening, and shading curtains. As mentioned above,

    this study focuses on natural ventilation in summer cooling, in conditions when

  • 8/16/2019 Benni 2016

    9/46

    8

    sun radiation can be neglected, therefore no operation of HVAC systems and

    shading curtains is considered during the trials and simulations. Indoor

    temperature and relative humidity are monitored in each sector with 10 min

    acquisition time; outdoor meteorological data of illuminance, temperature, wind

    speed and direction are measured with the same frequency. The material

    properties considered in the numerical models are reported in Table 1.

    2.2.  Computational approach

    Numerical modelling of airflows and temperature distributions was carried out

    through Autodesk CFD 2015 software, adopting a finite element approach

    [23][24]. This software uses several variants of the streamline upwind

    discretization schemes to model the advection terms and a true pressure-

    velocity algorithm based on the SIMPLE-R technique [25]. The finite element

    method is used to reduce the governing partial differential equations to a set of

    algebraic equations. In this method, the dependent variables are represented by

    polynomial shape functions over the element (a small area or volume). These

    representations are replaced into the governing partial differential equations and

    then the weighted integral of these equations over the element is taken where

    the weight function is chosen to be the same as the shape function. The result is

    a set of algebraic equations for the dependent variable at nodes on every

    element.

    2.2.1.  Governing equations

    The governing equations for fluid flow and heat transfer are the continuity

    equation, the momentum equations and the energy equation. The first is

    expressed as follows:

     

     

     

      0 

    (1)

  • 8/16/2019 Benni 2016

    10/46

    9

    where     is the density, t is time and u, v , w  are the velocity components along

    the global coordinate axis respectively x , y , z.

    In turbulence modelling, the software solves the time-averaged governing

    equations, which are obtained by assuming that the dependent variables can be

    represented as a superposition of a mean value and a fluctuating value about the

    mean. Using the notation that capital letters represent the mean values and

    lower case letters represent fluctuating values for the velocity components and

    pressure, the Reynolds averaged momentum equation referred to a generic  x  

    reference axis can be written as follows:

                    

      [2 ]

      [ (

     

    )]

      [ (

     

     )] 

    (2)

    where g x  is the gravitational acceleration in x  direction, P represents pressure,   

    is the fluid viscosity and  t  is the eddy viscosity. SDR is the distributed resistance

    term, which can be written in general as:

      (    )

    2    (3) 

    Where i   refers to the global coordinate direction, V i   represents the

    corresponding fluid velocity component, K   is a factor that can be determined

    from measurements of pressure drop versus flow rate,  f is the friction factor, DH 

    is the hydraulic diameter and C  is the viscosity coefficient, i.e. the inverse of the

    permeability.

    S  is the source term for rotating flow, which can be written in general as:

      2   (4)

  • 8/16/2019 Benni 2016

    11/46

    10

    where   is the rotational speed and r  is the distance from the axis of rotation.

    For incompressible and subsonic compressible flow, the Reynolds averaged

    energy equation, written in terms of static temperature T , is:

       

     

      

      [ ]

      [

    ]

      [

    ]  

    (5)

    Where C  p il the constant pressure specific heat, k  is the thermal conductivity, k t  

    the eddy conductivity and qV  is the volumetric heat source.

    The Mixing Length turbulence model was adopted, as it is primarily designed for

    internal natural convection analyses for gas flows. Specifically, the eddy viscosity

    is calculated according to the expression:

      √  

    (6)

    Where l m  is a mixing length calculated in the turbulent boundary layer as the

    product of the wall parameter and the normal distance from the wall, while the

    parameter G is given by:

    2 ()

    2 ()

    2 ( )

     

    (  ) (   ) (  ) 

    (7)

    2.2.2.   Advection scheme and convergence criteria

    For numerical stability, the advection terms are treated with upwind methods

    along with the weighted integral method. The advection scheme adopted is the

    Modified Petrov-Galerkin, a more stable variation of the Petrov-Galerkin one

  • 8/16/2019 Benni 2016

    12/46

    11

    [26]. It is suitable for all application types recommended with the latter, but

    typically produces more globally conservative results. In particular it has shown

    improvements as for accuracy of recirculating and secondary flows, natural

    convection stability, compressible flow accuracy and stability, and energy

    balance stability.

    Transient and steady state analyses were performed for model calibration. For

    the latter ones, convergence was assumed based on the union of the following

    criteria:

      the maximum instantaneous slope of the convergence data of all of the

    dependent variables from one iteration to the next is below the set level

    of 10-4;

      the slope of the convergence data averaged over several iterations of

    minimum, maximum and mean values of all of the dependent variables

    are below 0.01;

      the derivative of the maximum time averaged convergence slope

    (concavity) falls below the pre-determined level of 0.01;

     

    the fluctuation of the dependent variable about the mean value shows a

    standard deviation below the set level of 10-5.

    2.2.3.  Numerical modelling

    A two-dimensional simulation approach was adopted, as it is considered a

    computationally beneficial assumption for the investigation of the transport

    phenomena especially at the middle section of a structure with side vents

    opened along the whole length [22]. Under these conditions, the 2D model

    proved to be an accurate simplification of the three-dimensional phenomenon.

    In particular, two-dimensional simulations are used in scientific literature to

    investigate the effects of the ventilation configuration of tunnel greenhouses

    with inlet flow direction perpendicular to the longitudinal axis of the greenhouse

    [8,22]. According to literature, a large 2D geometrical model domain, of 160 m

    long and 50 m high, including the greenhouse at its centre, was used (Figure 2

  • 8/16/2019 Benni 2016

    13/46

    12

    shows the greenhouse and a portion of the surrounding domain). This domain

    was proposed in order to allow the interaction between external and indoor

    airflows to be naturally developed [27].

    The greenhouse was modelled as an empty building, in order to obtain results

    independent on the quantity or type of plants inside the structure, and on their

    arrangement. The building envelope was modelled with finite elements with

    physical properties corresponding to the real glass panels. The internal walls are

    not airtight, therefore the presence of several small openings allowing air

    passage among the various spans was modelled as an equivalent permeability of

    the internal walls corresponding to 1% of the length, uniformly distributed. The

    domain was subdivided into a number of surfaces necessary to identify the

    portions characterized by different initial conditions, and to develop meshes with

    detail levels suitable for the complexity of the phenomena to be modelled in the

    different zones of the domain. The domain was discretized into a hybrid grid that

    combine triangular and quadrilateral element types thus providing maximum

    flexibility in matching mesh cells with the boundary surfaces. It is well known

    that this generally leads to both accurate solutions and better convergence forthe numerical solution methods [28]. In particular, unstructured meshes were

    adopted for the various domain portions to generate the numerical grids outside

    and inside the greenhouse, with higher density in critical zones of the flow

    subject to strong gradients. Grid quality was enhanced through the placement of

    three layers of quadrilateral elements in resolving boundaries near solid walls

    and desks. The total number of elements in most of the results that are

    presented here was 24776, with average dimension of mesh sizes for the indoor

    domain of 0.2 m. This value was considered appropriate also in scientific

    literature with reference to the same investigation subject [8,22]. Mesh

    independency was verified by performing several steady state analyses with

    increasingly finer meshes simulating the indoor air. Eight meshes, from about

    13000 elements up to more than 52000 were adopted and the results in terms of

    air speed and pressure are reported in Figure 3. Transient analysis simulations

    with a larger number of elements  –  110947  – were also performed, to further

  • 8/16/2019 Benni 2016

    14/46

    13

    verify the independency of the numerical results from the grid features adopted,

    thus verifying the mesh sensitivity, as indicated in literature [29].

    2.3.  Measurement trials

    Calibration of the numerical modelling was performed by means of an ad-hoc

    data collection campaign in controlled environmental condition, and comparison

    with corresponding results obtained through CFD simulation. High frequency and

    spatial resolution of surveys allowed to obtain detailed data for accurate

    calibration. Monitoring of temperature and air speed was performed on 10

    February 2015. All data were measured on the vertical cross section intersecting

    the midpoint of the greenhouse length. In particular, temperature was recorded

    at the centre of the span at heights of 0.25m, 0.85 m, 1.45 m, 2.05 m, and 2.65 m

    and on the cultivation benches (0.85m height) at distances from the SE wall of

    0.90 m, 2.50 m, 3.60 m, 4.90 m, 7.30 m. Indoor temperature data were first

    collected with vents closed; then the side vent was opened and air speed and

    temperature were measured in the inlet edge. Temperature was measured with

    Delta Ohm thermo-hygrometer HD 2101 with resolution of 0.1°C, accuracy of

    ±0.1°C, operating temperature of-5°C ÷ +50°C, working relative humidity range of

    0 ÷ 90% RH without condensation, total measured values storage of 38000

    samples; together with stand-alone-data-logger PCE-HT71 with resolution of

    0.1°C, accuracy of ±0.5 C, measurement range of −40 C ÷ +70 C, total memory

    storage of 16000 records. Air speed was measured through Delta Ohm DO9847

    multi-function datalogger, with working temperature range from -5°C ÷ +50°C,

    RH range 0 ÷ 90% not condensing, storage temperature range from -20°C ÷

    +60°C. Vane probe  anemometer adopted was model AP472 S1, with 100 mm

    diameter, speed measurement range from 0.4 m/s ÷ 30 m/s, resolution of 0.01

    m/s, accuracy of ± (0.1 m/s + 1.5% full scale). Hot wire probe anemometer

    adopted was model AP471 S4, with measurement range from 0.1 m/s ÷ 5 m/s,

    resolution of 0.01 m/s, accuracy of ±0.05 m/s (from 0 to 0.99 m/s) and ±0.15 m/s

    (from 1.00 m/s to 5.00 m/s). Data were collected with constant time step of 5s.

  • 8/16/2019 Benni 2016

    15/46

    14

    Boundary and initial conditions of the model (Table 2) were defined on the basis

    of the measured data. In particular, inlet air speed in the domain was defined

    based on meteorological data recorded, while outdoor air temperature was set

    on the basis of data measured through the external sensors. A symmetry

    condition was applied at the top of the domain, while at the bottom boundary

    ground temperature of the greenhouse was set at a value obtained from

    temperature measured close to the floor during the trials. Differences between

    the floor temperature and the air temperature in the position of the indoor

    sensor were recorded and their average value was adopted to assess the ground

    temperature for simulations under conditions when only data of indoor air

    temperature were available. These estimates were adopted as boundary

    conditions for numerical simulations.

    Air properties have been defined constant on the basis of CFD preliminary

    analyses carried out with transient approach referred to the conditions of Table

    2, considering either fixed or variable material properties.

    Measured and simulated data were compared for validation by computing the

    RMSE, defined according to the usual expression:

    RMSE= ∑   −̂^=1   , (8)

    where N is the number of observations, yi is the measured parameter, and ŷ is

    the predicted value.

    Moreover, the ratio of percentage deviation (RPD) index was considered for the

    time histories of the parameters analysed. RPD is defined as the ratio of standard

    deviation of measured data to RMSE of predicted values. The model was

    assessed according to Rossel et al. [30]: RPD < 1 means “very poor” model; 1 ≤

    RPD < 1.4 means “poor” model; 1.4 ≤ RPD < 1.8 means the model is “fair”,

    therefore it can be adopted for general considerations; 1.8 ≤ RPD < 2 indicates a

  • 8/16/2019 Benni 2016

    16/46

    15

    “good” model; 2 ≤ RPD < 2.5 indicates a “very good” model; and RPD ≥ 2.5 means

    the model is “excellent”. 

    The soundness of our model regarding the prediction on air speed was further

    assessed according to scientific literature [31], by comparing the RMSE of our

    calibration with that obtained from a statistical model adopted as a benchmark

    for short-term predictions of air speed. The statistical model adopted is a multi-

    linear regression one (MLR) developed by Mathur and Mathur [32], where the

    following quantities are the independent variables:

    -  weighted moving average (MA) defined as follows 

    MA2, t  y1∗p1+y∗pp1+p   (9) where yt-1  and yt-2  are air speed respectively at time (t -1) and (t -2); pi  are

    weights. Lag was choses equal to 2 based on autocorrelation lower limit of

    0.7. Autocorrelation values obtained were adopted as weights p1 and p2;

    -  Exponential moving average EMA(2,t ) with lag of 2 data, defined as

    EMA(2,1)=y 0; EMA(2,t )=    * y t- 1+(1-   )*EMA(2,t -1), with   

    =2/(lag+1)=2/3; 

    -  Oscillator (OSC), defined as the difference of moving averages or exponential

    moving averages of two different periods: OSC (t )= EMA(3,t )-EMA(2,t ), where

    EMA(3,t ) is defined as EMA(3,0)= y 0 ; EMA(3,t )=   *y t-1+(1- )*EMA(3,t -1),

    with  =2/(lag+1)= 1/2

    -  Rate of change within time steps ROC(t ); 

    -  Second and third moment of the series M2(t ) and M3(t ). The r th moment of y ,

    Mr  is given by 

       ∑ −̅   (10) where ̅ is the mean of N variables. 

    Therefore the autoregressive model is expressed as follows:

    ŷ (t ) = a0 + a1*MA(2,t )+a2*OSC (t ) + a3*ROC (t ) +

    +a4*EMA(2,t ) + a5*M2(t ) + a6*M3(t ) (11)

    2.4.  Numerical simulations

  • 8/16/2019 Benni 2016

    17/46

    16

    The study aims at investigating the effect of vents opening in terms of natural

    ventilation, independently on the other factors potentially affecting indoor

    climate, in particular solar radiation. Therefore the analyses were performed in

    condition where solar radiance can be neglected, namely for hour angles of 80°

    or more. In this regard the study by [22] found that airflows and the temperature

    distribution in a greenhouse are significantly altered by the consideration of

    external incident radiation in the calculations when hour angles are between -60°

    and +60° (being noon 0°).

    The greenhouse system of automatic vent opening of the study case is

    programmed to fully open all the windows of a span when indoor air

    temperature overcomes a threshold value, usually set at 26°C. Numerical

    simulations were performed to assess the efficacy of this solution in comparison

    with the other possible vents opening configurations. The conditions recorded on

    8 August 2014 since 19.30 to 20.30 were considered as the study case for

    simulating the alternative scenarios and assessing their performance in summer

    cooling. Indoor temperature and meteorological conditions recorded are

    reported in Table 3. Wind direction was from SE. The resulting boundary andinitial conditions are reported in Table 4.

    Simulations were performed to analyse the following scenarios of vent opening

    configuration, summarized in Table 5. The scenario AO (All Open) was defined as

    the real operating condition of the greenhouse span, i.e. with all the vents open

    during the considered time interval; scenario WC (Windward Closed) was defined

    as the condition of side vent open, roof windward vent closed and roof leeward

    vent open; the scenario LC (Leeward Closed) was defined as the condition with

    side vent open, windward roof vent open and leeward roof vent closed.

    3.  Results and Discussion

  • 8/16/2019 Benni 2016

    18/46

    17

    3.1.  Calibration

    Both steady state and transient CFD simulations of the greenhouse were

    performed under the environmental conditions surveyed during the calibration

    trials.

    Steady state analyses were developed to simulate the conditions surveyed in the

    calibration trial after 420 s since wind opening: from that moment, both air

    speed and air temperature showed a steady behaviour for about 140 s (Figure 4).

    Thus air speed results of the steady state simulation were compared with the

    corresponding average data measured over that period of 140 s (28 data).

    Average air speed measured was 0.37 m/s, and simulated one in the same point

    in steady state conditions resulted 0.36 m/s. Therefore the model proved to

    represent real data with excellent soundness.

    Transient analyses results regarding air speed and temperature were then

    compared to measured data (Figure 5a).

    As measured data of air speed showed high-frequency oscillations with low-

    amplitude, data were processed according to Hilbert-Huang transformations: six

    intrinsic mode functions were found for measured data and four for simulated

    data. The intrinsic mode function with lowest amplitude of measured data was

    omitted, thus obtaining the smoothed trend of Figure 5b: the corresponding

    RMSE is 0.0559 m/s, which is very close to anemometer accuracy and therefore

    indicates a good agreement between measured and simulated data.

    A further quantification of the agreement between measured and simulated air

    speed data was performed by developing an autoregressive model, as illustrated

  • 8/16/2019 Benni 2016

    19/46

    18

    in the methodology section. The model was calibrated based on the first 100

    time steps and was applied for an interval corresponding to half the calibration

    period. RMSE between measured data and the autoregressive model was 0.0967

    m/s; RMSE between measured data and CFD simulated data was 0.0974 m/s.

    Therefore the CFD model proved as sound as the statistical predictive model

    based on the measured data.

    Air temperature measured in the point of anemometric survey was compared

    with corresponding CFD simulated results (Figure 6). Resulting RPD is 1.6091,

    thus qualifying fair soundness of the model.

    Temperature trends monitored inside the greenhouse were compared with ones

    simulated through transient analyses. Data referred to the centre of the span

    width at 0.85 m and 2.05 m height showed good agreement (Figure 7) between

    surveyed and computed values: in fact RPD on the average trend is 1.721 (fair

    model).

    3.2.  Simulations

    As mentioned above, once the model had been validated, it was applied to

    analyse the effects of different vents opening configurations on temperature

    reduction in hot climate conditions. The simulated temperatures for the scenario

    AO were compared with temperature data recorded by the standard greenhouse

    management system, with reference to the point where the thermometer of the

    greenhouse facility is installed. The comparison provided a RMSE equal to 0.32°C,

    thus allowing a further confirmation of the reliability of the model.

  • 8/16/2019 Benni 2016

    20/46

    19

    Simulated temperature and air speed data were computed above the centre of

    each workbench at heights of 1.00 m, 1.50 m, 2.00 m,2.50 m, thus obtaining a

    grid with 4 rows and 3 columns. Such grid is representative of the microclimatic

    condition of the indoor working and cultivation area. Mean temperature data

    trends (over these twelve points) of the three opening configurations are

    reported in Figure 8. These temperature trends resulted substantially similar

    among rows and columns of the grid.

    Scenario WC proved to lead to the greatest temperature reduction, although the

    scenario with leeward vents closed caused a faster initial temperature drop. In LC

    scenario, after a first period of about 20 min of circulation of outside air in the

    cultivation area, the prevailing airflow becomes almost steady and consists in

    direct passage from the side wall vent to the leeward roof vent, with poor

    contribution to temperature reduction in the cultivation area. The reduction of

    outside air speed characterizing the final part of the simulation period causes

    only slight temperature decrease. Scenarios AO and WC show a very similar

    temperature reduction path, but their effects become opposite when outside

    wind speed becomes near to zero, after 35 min from the beginning of the

    simulation. The scenario LC entails further cooling, while temperatures in WC

    scenario do not decrease anymore. Therefore, in presence of outside wind, the

    opening of wall vent on the windward side together with roof leeward vent

    proves more effective than the opening of both roof vents, to achieve air

    exchange in the cultivation area. In fact, as it is shown in Figures 9 and 10, AO

    conditions cause the formation of a direct airflow through the roof openings,

    while the scenario WC enhances indoor air circulation besides relevant air

    exchange between outside and inside; the scenario LC is clearly unfavourable to

    outbound airflows.

  • 8/16/2019 Benni 2016

    21/46

    20

    Temperature profiles along the span width at 1.50 m above the ground for the

    three scenarios, at 15 min, 30 min, and 60 min are reported in Figure 11.

    Average temperature reductions over the whole simulation period are 3.37°C

    with WC, 2.77°C with LC, and 2.54°C with AO. The performances of the three

    scenarios were assessed on the basis of a linear scale where 0 indicates no

    reduction of temperature over the simulation period, and 100% represents a

    sudden reduction of indoor temperature to outdoor one since the first time step

    (average reduction of 5.23°C). Resulting performances are 64% for WC, 53% for

    LC, and 48% for AO. Such percentages give an indication of the ratio of heat

    removed from inside the greenhouse in the three scenarios (Figure 12). While

    the effects of AO and LC scenarios are very close to each other, the performance

    of WC solution entails a significantly greater cooling.

    The trends of mean air speed are reported in Figure 13, and for each scenario

    they are substantially similar among rows and columns of the grid.

    Indoor air speed in the three scenarios resulted always far below the

    acceptability limit for proper working conditions in greenhouses [2]. The scenario

    WC entails higher rates of air speed for almost the whole duration of the

    simulation, except for a couple of minutes when outside wind decreases to zero:

    in this condition, indoor air speed becomes equal for the three scenarios.

    Mean modules of the air speed components perpendicular to each inlet edge

    was then computed along the edges themselves at every time step. Results are

  • 8/16/2019 Benni 2016

    22/46

    21

    reported in Figure 14 , where positive values represent inlet flows, negative

    outlets.

    Overall average airflows through the windows, referred to one meter of span

    length (inlet section width is 0.70 m for all vents), are reported in Table 6.

    It should be noted that airflow balance for each scenario is different from zero,

    because of air exchanges with the other spans, considered through proper

    definition of filtration rates in internal walls. Table 6 clearly shows that the

    scenario LC hampers outbound airflows and entails overall air inlet (54.58

    m3/h/m) less than half of AO (121.18 m3/h/m) and also smaller than WC. This

    phenomenon appears to be the main reason of the negative performance of

    such scenario in terms of temperature reduction.

    4. 

    Conclusions

    The study analysed the effects of roof and side vents opening configuration on

    temperature reduction in a study case of glass greenhouse in hot climatic

    conditions. A finite element CFD model of the greenhouse was developed and

    calibrated through experimental monitoring of indoor and outdoor climatic

    parameters. Numerical simulations of indoor conditions under real outdoor

    climatic parameters recorded in a selected time interval of over one hour wereperformed. Various possible configurations of opening and closure of roof vents

    were set, with a side wall vent always open. The analyses also accounted for the

    outside air speed variable, including the case of absence of wind. The results

    were assessed in terms of effectiveness in indoor cooling. Indoor temperature

    distributions and trends were analysed, together with indoor air speed and

    airflows. The best performances were obtained with windward vent closed,

    which entailed 64% of the maximum heat removal achievable through natural

  • 8/16/2019 Benni 2016

    23/46

    22

    ventilation. The other scenarios considered showed a performance index of

    about 50%.

    The automatic system of vents management of the greenhouse, which is a

    commercial and widespread solution, is set to open all roof vents when

    temperature overcomes a predefined threshold. This proved significantly less

    effective than the scenario where windward roof vents were closed. Therefore,

    the results suggest to enhance the vents control system by considering also wind

    direction as input, so that the windward roof-vents closed scenario could be

    adopted for the upwind span of the greenhouse. The results represents the basis

    of further developments, which are ongoing and are expected to include the

    contribution of solar radiation and analyse how the presence of plants can

    modify the results obtained in this study. The outputs are expected to complete

    the framework of programming strategies to enhance the automatic windows

    control devices of greenhouses.

    Acknowledgements

    The authors wish to thank prof. Beatrice Pulvirenti PhD, Assistant Professor at

    the Department of Industrial Engineering of the University of Bologna, for the

    support provided in CFD modelling.

    References

    [1] N. Katsoulas, T. Bartzanas, T. Boulard, M. Mermier, C. Kittas, Effect of

    Vent Openings and Insect Screens on Greenhouse Ventilation, Biosyst.

    Eng. 93 (2006) 427 –436. doi:10.1016/j.biosystemseng.2005.01.001.

    [2] C.A. Campotti, A. Latini, S. Matteo, C. Viola, CERTIFICATI BIANCHI Guida

    Operativa per la Scheda 40E - Gli incentivi del DM 28 dicembre 2012 sui

    sistemi serra, ENEA, Rome, 2014.

  • 8/16/2019 Benni 2016

    24/46

    23

    [3] N. Bennis, J. Duplaix, G. Enéa, M. Haloua, H. Youlal, Greenhouse climate

    modelling and robust control, Comput. Electron. Agric. 61 (2008) 96 –107.

    doi:10.1016/j.compag.2007.09.014.

    [4] E. Fitz-Rodríguez, C. Kubota, G.A. Giacomelli, M.E. Tignor, S.B. Wilson, M.

    McMahon, Dynamic modeling and simulation of greenhouse

    environments under several scenarios: A web-based application, Comput.

    Electron. Agric. 70 (2010) 105 –116. doi:10.1016/j.compag.2009.09.010.

    [5] K.S. Kumar, K.N. Tiwari, M.K. Jha, Design and technology for greenhouse

    cooling in tropical and subtropical regions: A review, Energy Build. 41

    (2009) 1269 –1275. doi:10.1016/j.enbuild.2009.08.003.

    [6] K.S. Kumar, M.K. Jha, K.N. Tiwari, A. Singh, Modeling and evaluation of

    greenhouse for floriculture in subtropics, Energy Build. 42 (2010) 1075 –

    1083. doi:10.1016/j.enbuild.2010.01.021.

    [7] F.D. Molina-Aiz, D.L. Valera, A.A. Peña, J.A. Gil, A. López, A study of

    natural ventilation in an Almería-type greenhouse with insect screens by

    means of tri-sonic anemometry, Biosyst. Eng. 104 (2009) 224 –242.

    doi:10.1016/j.biosystemseng.2009.06.013.

    [8] T. Bartzanas, T. Boulard, C. Kittas, Effect of Vent Arrangement on

    Windward Ventilation of a Tunnel Greenhouse, Biosyst. Eng. 88 (2004)

    479 –490. doi:10.1016/j.biosystemseng.2003.10.006.

    [9] A. Ganguly, S. Ghosh, Model development and experimental validation of

    a floriculture greenhouse under natural ventilation, Energy Build. 41

    (2009) 521 –527. doi:10.1016/j.enbuild.2008.11.021.

    [10] I.M. Al-Helal, S.A. Waheeb, A.A. Ibrahim, M.R. Shady, A.M. Abdel-Ghany,

    Modified thermal model to predict the natural ventilation of greenhouses,

    Energy Build. 99 (2015) 1 –8. doi:10.1016/j.enbuild.2015.04.013.

    [11] C. Kittas, T. Boulard, T. Bartzanas, N. Katsoulas, M. Mermier, Influence of

    an insect screen on greenhouse ventilation, Trans. Am. Soc. Agric. Eng. 45

    (2002) 1083 –1090.

    [12] H. Demrati, T. Boulard, A. Bekkaoui, L. Bouirden, SE—Structures and

  • 8/16/2019 Benni 2016

    25/46

    24

    Environment, J. Agric. Eng. Res. 80 (2001) 261 –271.

    doi:10.1006/jaer.2001.0740.

    [13] C. Kittas, N. Katsoulas, T. Bartzanas, M. Mermier, T. Boulard, The impact

    of insect screens and ventilation openings on the greenhouse

    microclimate, Trans. ASABE. 51 (2008) 2151 –2165.

    [14] F.D. Molina-Aiz, H. Fatnassi, T. Boulard, J.C. Roy, D.L. Valera, Comparison

    of finite element and finite volume methods for simulation of natural

    ventilation in greenhouses, Comput. Electron. Agric. 72 (2010) 69 –86.

    doi:10.1016/j.compag.2010.03.002.

    [15] F. Ayuga, Present and future of the numerical methods in buildings andinfrastructures areas of biosystems engineering, J. Agric. Eng. 46 (2015) 1.

    doi:10.4081/jae.2015.436.

    [16] M. Teitel, O. Liran, J. Tanny, M. Barak, Wind driven ventilation of a mono-

    span greenhouse with a rose crop and continuous screened side vents and

    its effect on flow patterns and microclimate, Biosyst. Eng. 101 (2008) 111 –

    122. doi:10.1016/j.biosystemseng.2008.05.012.

    [17] M. Teitel, E.J. Baeza, J.I. Montero, Greenhouse design: Concepts and

    trends, in: Acta Hortic., 2012: pp. 605 –620.

    [18] M. Teitel, E. Wenger, Air exchange and ventilation efficiencies of a

    monospan greenhouse with one inflow and one outflow through

    longitudinal side openings, Biosyst. Eng. 119 (2014) 98 –107.

    doi:10.1016/j.biosystemseng.2013.11.001.

    [19] I.-B. Lee, J.P.P. Bitog, S.-W. Hong, I.-H. Seo, K.-S. Kwon, T. Bartzanas, et al.,

    The past, present and future of CFD for agro-environmental applications,

    Comput. Electron. Agric. 93 (2013) 168 –183.

    doi:10.1016/j.compag.2012.09.006.

    [20] D.K. Fidaros, C.A. Baxevanou, T. Bartzanas, C. Kittas, Numerical simulation

    of thermal behavior of a ventilated arc greenhouse during a solar day,

    Renew. Energy. 35 (2010) 1380 –1386. doi:10.1016/j.renene.2009.11.013.

    [21] http://www.climatedata.eu/climate.php?loc=itxx0006&lang=en, (2016).

  • 8/16/2019 Benni 2016

    26/46

    25

    [22] C. Baxevanou, D. Fidaros, T. Bartzanas, C. Kittas, Numerical simulation of

    solar radiation, air flow and temperature distribution in a naturally

    ventilated tunnel greenhouse, Agric. Eng. Int. CIGR J. 12 (2010) 48 –67.

    [23] J.G. Rice, R.J. Schnipke, An equal-order velocity-pressure formulation that

    does not exhibit spurious pressure modes, Comput. Methods Appl. Mech.

    Eng. 58 (1986) 135 –149. doi:10.1016/0045-7825(86)90093-9.

    [24] R.J. Schnipke, J.G. Rice, FINITE ELEMENT METHOD FOR FREE AND FORCED

    CONVECTION HEAT TRANSFER., Int. J. Numer. Methods Eng. 24 (1987)

    117 –128.

    [25] S. Patankar, Numerical heat transfer and fluid flow, in: Ser. CoputationalMethods Mech. Therm. Sci., 1980: pp. 1 –197.

    [26] C. de Falco, E. O’Riordan, A parameter robust Petrov–Galerkin scheme for

    advection –diffusion –reaction equations, Numer. Algorithms. 56 (2010)

    107 –127. doi:10.1007/s11075-010-9376-y.

    [27] E. Rico-García, Aerodynamic study of greenhouses using computational

    fluid dynamics, Int. J. Phys. Sci. 6 (2011). doi:10.5897/IJPS11.852.

    [28] J. Tu, G.-H. Yeoh, C. Liu, Computational Fluid Dynamics, 2013.

    doi:10.1016/B978-0-08-098243-4.09994-7.

    [29] S. Di Sabatino, R. Buccolieri, B. Pulvirenti, R. Britter, Simulations of

    pollutant dispersion within idealised urban-type geometries with CFD and

    integral models, Atmos. Environ. 41 (2007) 8316 –8329.

    doi:10.1016/j.atmosenv.2007.06.052.

    [30] R.A. V Rossel, H.J. Taylor, A.B. McBratney, Multivariate calibration of

    hyperspectral ??-ray energy spectra for proximal soil sensing, Eur. J. Soil

    Sci. 58 (2007) 343 –353. doi:10.1111/j.1365-2389.2006.00859.x.

    [31] J. Jung, R.P. Broadwater, Current status and future advances for wind

    speed and power forecasting, Renew. Sustain. Energy Rev. 31 (2014) 762 –

    777. doi:10.1016/j.rser.2013.12.054.

    [32] P. Mathur, S. Mathur, A Simple Weather Forecasting Model Using

  • 8/16/2019 Benni 2016

    27/46

    26

    Mathematical Regression, Indian Res. J. Ext. Educ. I (2012) 161 –168.

  • 8/16/2019 Benni 2016

    28/46

    27

    Figure 1. Cross section of the greenhouse.

  • 8/16/2019 Benni 2016

    29/46

    28

    Figure 2. Central portion of the FEM mesh with temperature contour of initial

    condition for model calibration.

  • 8/16/2019 Benni 2016

    30/46

    29

    Figure 3. Mesh convergence diagram in terms of air speed (v ) computed near

    inlet opening and pressure difference ( p) between points at heights of 4 m and

    1m in the middle of the greenhouse span.

    0%

    20%

    40%

    60%

    80%

    100%

    120%

    0 10000 20000 30000 40000 50000 60000

    nr. of elements

    Mesh convergence

     p

  • 8/16/2019 Benni 2016

    31/46

    30

    Figure 4. Trends of air speed and temperature in the observation period selected

    for steady state CFD calibration analyses.

  • 8/16/2019 Benni 2016

    32/46

    31

    a)

    b)  

    Figure 5. a) Trends of air speed measured near the side wall vent and

    corresponding data obtained though CFD transient simulation; b) comparison of

    the same data processed through Hilbert-Huang transformation to remove the

    high-frequency low-amplitude oscillations.

  • 8/16/2019 Benni 2016

    33/46

    32

    Figure 6.Trend of air temperature in the air speed measurement point.

  • 8/16/2019 Benni 2016

    34/46

    33

    Figure 7. Comparison between measured and simulated temperature at different

    heights in the centre of the greenhouse. The continuous line indicates identity,

    the dotted ones indicate differences of 0.5°C, the dash-dotted ones differences

    of 1°C.

  • 8/16/2019 Benni 2016

    35/46

    34

    Figure 8. Mean temperature trends in the cultivation area for the three opening

    configuration considered.

  • 8/16/2019 Benni 2016

    36/46

    35

    Figure 9 Temperature contour and air velocity vectors in transient analysis for

    the three scenarios at 20 s.

    AO WC  LC 

    T(°C) 

  • 8/16/2019 Benni 2016

    37/46

    36

    Figure 10 Temperature contour and air velocity vectors in transient analysis for

    the three scenarios at 2400 s.

    AO WC  LC 

    T(°C) 

  • 8/16/2019 Benni 2016

    38/46

    37

    Figure 11. Horizontal temperature profile at height of 1.50m for the three

    scenario at time steps of 15 min, 30 min, 60 min.

  • 8/16/2019 Benni 2016

    39/46

    38

    Figure 12. Performances of the three scenarios, in terms of heat removal form

    the greenhouse.

  • 8/16/2019 Benni 2016

    40/46

    39

    Figure 13 Mean air speed in the cultivation area.

  • 8/16/2019 Benni 2016

    41/46

    40

    Figure 14. Air speed trends through vents in the various scenarios.

  • 8/16/2019 Benni 2016

    42/46

    41

    Table 1. Material properties.

    Material Density (kgm-3) Viscosity (Pa s) Conductivity

    (Wm-1K-1)

    Specific Heat

    (J kg-1 K-1)

    Air 1.20473 1.817 0.02563 1004

    Glass 2700 - 0.78 840

  • 8/16/2019 Benni 2016

    43/46

    42

    Table 2. Boundary conditions and initial conditions of CFD analyses for calibration

    (10 February 2015).

    Boundary conditions Initial conditions

    Outdoor air

    speed

    Floor

    temperature

    Outdoor

    temperature

    Indoor temperature

    0.1 m/s 22°C 9 °C

    Height (m) Temperature

    (°C)

    < 0.4 18

    0.4-0.8 20

    0.8-1.4 23.5

    1.4-1.8 22

    1.8-2.6 23.7

    2.6-3.6 28.7

    > 3.6 30

  • 8/16/2019 Benni 2016

    44/46

    43

    Table 3. Indoor and outdoor climatic parameters recorded during the reference

    period considered for CFD simulations.

    Date, time Indoor

    T (°C)

    Outdoor

    T (°C)

    Wind speed

    (ms-1)

    08/08/2014, 19.30 27.6 25.2 0.67

    08/08/2014, 19.40 27.6 25.2 0.89

    08/08/2014, 19.50 27.2 24.8 0.67

    08/08/2014, 20.00 27.2 24.8 0.67

    08/08/2014, 20.10 26.8 24.4 008/08/2014, 20.20 26.8 24.4 0.11

    08/08/2014, 20.30 26.8 24.0 0

  • 8/16/2019 Benni 2016

    45/46

    44

    Table 4. Boundary conditions and initial conditions of CFD analyses for

    simulations (8 August 2014).

    Boundary conditions Initial conditions

    Outdoor air

    speed

    Floor

    temperature

    Outdoor

    temperature

    Indoor temperature

    time

    (s)

    v

    (m/s)

    22°C 25.2°C

    Height (m) Temperature

    (°C)

    0 0.67 < 0.4 22

    600 0.89 0.4-0.8 25.5

    1200 0.67 0.8-1.4 27.6

    1800 0.67 1.4-1.8 29.5

    2400 0 1.8-2.6 30.5

    3000 0.11 2.6-3.6 31.5

    3600 0 > 3.6 32.5

  • 8/16/2019 Benni 2016

    46/46

    Table 5. Combinations of vent opening and closure for the scenarios considered

    in CFD simulations.

    Scenario AO Scenario WC Scenario LC

    Side vent Open Open Open

    Roof Windward vent Open Closed Open

    Roof Leeward vent Open Open Closed

    Table 6. Airflows through greenhouse windows in the analysed scenarios.

    Airflow (m3/h/m) Side wall vent Windward roof vent Leeward roof vent

    Scenario AO 61.73 59.45 -77.50

    Scenario WC 73.47 0 -50.62

    Scenario LC 26.15 28.43 0