Final

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Heat Transfer Research 49(6):509–528 (2018)

NUMERICAL STUDY OF A SOLAR GREENHOUSE DRYER


WITH A PHASE-CHANGE MATERIAL AS AN ENERGY
STORAGE MEDIUM

Orawan Aumporn,1 Belkacem Zeghmati,1,* Xavier Chesneau,1


& Serm Janjai2
1
Laboratoire de Mathématiques et PhySique (LAMPS), Université de Perpignan Via Domitia
 52, Avenue Paul Alduy, F-66860 Perpignan cedex, France
2
Solar Energy Research Laboratory (SERL), Department of Physics, Faculty of Science, Silpakorn
 University, Nakhon Pathom 73000, Thailand
*Address all correspondence to: Belkacem Zeghmati, Laboratoire de Mathématiques et PhySique
(LAMPS), Université de Perpignan Via Domitia 52, Avenue Paul Alduy, F-66860 Perpignan cedex,
 France, E-mail: zeghmati@univ-perp.fr

Original Manuscript Submitted: 3/17/2017; Final Draft Received: 7/7/2017

A numerical study of the thermal behavior of a solar greenhouse dryer with a thermal energy storage unit is presented.
The solar greenhouse dryer consists of a gothic metallic arch structure covered with a polycarbonate film on a metallic plate
floor. The products to be dried (100 kg of banana Musa ABB CV. Kluai "Namwa") are located as a thin layer on four me-
tallic grids. The thermal energy storage unit, disposed under the greenhouse floor, is composed of a layer of phase-change
materials (PCM) placed between the metal plate and a concrete slab. Paraffin wax was used as PCM in thermal energy
storage with a melting temperature of 28oC. Transfer equations are derived by considering energy balance for different
components of the greenhouse dryer. The enthalpy method and heat conduction equation have been used for calculating the
PCM layer and concrete slab, respectively. Equations are solved numerically by an implicit finite difference scheme and
homemade software. Parametric studies of the greenhouse dryer coupled to the thermal energy storage unit illustrate the
effects of drying air volume flow rate on the greenhouse temperature, drying duration, as well as the efficiency of the solar
dryer and energy storage unit. Our results allowed us to conclude that under the climatic conditions of Nakorn Pathom
(Thailand) the thermal storage unit improves the greenhouse solar dryer efficiency. For instance, for the drying air volume
flow rate ranging from 0.05 to 0.2 m3·s–1, this efficiency varies between 12% and 38% with a thermal storage unit and
between 8% and 28% without a storage unit.

KEY WORDS: phase-change material, modeling, greenhouse dryer, solar dryer efficiency, storage efficiency

1. INTRODUCTION
In recent years, most developing countries were facing the problem of energy crisis because of the large gap be-
tween demand and supply of energy. This problem can be minimized by utilizing renewable energy such as solar
energy and an energy storage unit. Due to the intermittence of solar radiation, energy storage systems are used to
store excess energy during the peak time of solar radiation to be used during off-sun hours or when the amount
of available energy is inadequate. So, the use of solar energy in the process of drying agricultural products has
become one of the most popular and attracting economic investments.
Solar dryers integrated with a thermal storage system can solve the problems of interrupted solar irradiation and
peak temperature rise at noon (Shalaby and Bek, 2014). The thermal energy can be stored in the form of sensible
heat, latent heat or chemical energy (Bal et al., 2011). In the sensible heat storage, thermal energy is stored either

1064-2285/18/$35.00 © 2018 by Begell House, Inc. www.begellhouse.com 509


510 Aumporn et al.

NOMENCLATURE
2
A area, m ciel sky
–1 –1
Cp specific heat, J·kg ·K cond conduction
–1
DB dry basis, kg water (kg dry matter) conv convection
e thickness, m conv, a-f convection between drying air
f liquid fraction and metal plate
–1
H total volumetric enthalpy, J·kg conv, a-pr convection between drying air
–2 –1
h heat transfer coefficient, W·m ·K and product
HL sunrise time, h conv, c-a convection between cover and
HR air relative humidity, % drying air
–1 –1
k thermal conductivity, W·m ·K conv, c-amb convection between cover
–1
L latent heat, J·kg and ambient
Lo length of the greenhouse dryer, m d retrieval
m mass, kg f metal plate
P pressure, Pa fus fusion
Pa atmospheric pressure, Pa g greenhouse
Pu useful power, W init initial
3 –1
Qv air volume flow rate, m ·s max maximum
Re Reynolds number mcp phase-change material
Sc Schmidt number min minimum
T temperature, K pr product
t time, s r radiative
–1
v air velocity, m·s r, c-ciel radiation between cover and sky
Z height of heat storage system r, c-f radiation between cover and metal
plate
Greek Symbols
r, c-se radiation between cover and floor outside
α absorptivity the greenhouse dryer
η efficiency r, pr-c radiation between product and cover
–3
ρ density, kg·m r, pr-f radiation between product and metal
τ transmittivity plate
–2
φ solar irradiance, W·m s storage
se floor outside the greenhouse
Subscripts
dryer
a drying air Terre soil
amb ambient v water vapor
b concrete vap vaporization
c cover vs saturation water vapor

Heat Transfer Research


Solar Greenhouse Dryer with a Phase-Change Material as an Energy Storage Medium 511

in solid materials (e.g., sand, gravel, and bricks) or in a liquid (water). The solid and liquid used as sensible heat
storage materials for drying agricultural products were considered by Bal et al. (2010). However, the latent heat
storage is a more attractive form because of the high energy storage capacity, absorbing and releasing heat at
almost a constant temperature, chemical stability, and small volume change during phase transformation (Abhat,
1983; Duffie and Beckman, 2013). Suitable phase-change materials (PCMs) include paraffin waxes, fatty acids,
hydrated salts, calcium chloride hexahydrate, sodium thiosulfate pentahydrate, and sodium carbonate decahydrate
(Zalba et al., 2003) and as latent heat use thermal energy storage; applications are reviewed in Abhat (1983) and
Zalba et al. (2003).
Various experimental and numerical studies have been reported in the literature to evaluate the PCMs thermal
energy in solar dryers. A detailed review of PCMs used in a thermal energy storage unit for a solar dyer was given
by Shalaby et al. (2014), Bal et al. (2010), and Sharma et al. (2009). Among these studies, an indirect solar dryer
using PCMs as an energy storage medium was experimentally investigated for drying medical plants (Shalaby
et al., 2014). The results of the study showed that using paraffin wax as PCM to store the solar energy, the tem-
perature of the drying air is to be higher than the ambient temperature by 2–7oC after sunset during five hours
at least for a wide range of air mass flow rates (0.0664–0.2182 kg·s–1). It is possible to dry agricultural products
in the temperature range 40–75oC in solar dryers with a heat storage unit with PCM (paraffin wax) (Bal et al.,
2011). A hybrid solar dryer with a heat storage unit composed of paraffin wax as PCM was investigated exper-
imentally by Reyes et al. (2014) for drying mushrooms. It was indicated that the thermal efficiency depending on
the meteorological conditions fluctuated between 22% and 62%, while the efficiency of the accumulator panel varied
between 10% and 21%. The energy accumulated in paraffin wax extended the drying process at least for two hours
and reduced the electrical energy consumption for drying by 40–70%. Devahastin and Pitaksuriyarat (2006) studied
the effect of using paraffin wax (melting temperature 54οC) in a thermal energy storage unit for drying food product
in an indirect solar dryer. The results presented showed that the energy storage unit was able to reduce the required
thermal energy by about one third to dry sweet potato with the inlet air velocity varying in the range 1–2 m·s–1. It
was additionally reported that for the inlet drying air temperature between 70oC and 90oC, the amount of the energy
supplied by the thermal energy storage was 1920 and 1380 kJ·min–1·kg–1 for the drying air velocity equal to 1 and
2 m·s–1, respectively. The saved energy was about 40% and 34% corresponding to these velocities. The drying of
seeded grapes in a hybrid solar dryer with a thermal energy storage unit composed of PCM was investigated ex-
perimentally by Çakmak and and Yildiz (2011). Calcium chloride hexahydrate as PCM was disposed in the lower
section of one of the two solar collectors of a solar dryer. The results were focused on the relationship between the
air velocity increment and drying time period. Esakkimuthu et al. (2013) investigated experimentally an indirect
solar dryer with a thermal storage unit composed of PCM (HS 58 of melting temperature 58οC) to analyze the
charging and discharging periods of the storage unit. The results showed that this storage unit is able to provide
an air temperature equal to 50οC during the discharging period of 4 h for an air mass flow rate equal to 0.0555
kg·s–1. They concluded that the PCM thermal energy storage unit improves the thermal performance of the dryer.
Most of solar dryers have been developed for household needs or small-scale commercial applications. Howev-
er, in some agro-industrial applications, larger dryers as solar greenhouse ones are required. A number of studies
have been reported on crop drying in greenhouse dryers (Patil and Gawande, 2016; Bala et al., 2003; Srisittipoka-
kun et al., 2012; Hossain and Bala, 2007; Janjai et al., 2008; Kaewkiew et al., 2012; Palled et al., 2012; Dulawat
and Rathore, 2012; Kagande et al., 2012).
Many studies have been reported on solar greenhouse using PCMs as a thermal energy storage medi-
um (Lazaar et al., 2014; Kürklü et al., 1977; Boulard et al., 1990; Berroug et al., 2011; Kumari et al., 2006).

Volume 49, Issue 6, 2018


512 Aumporn et al.

Lazaar et al. (2014) experimentally studied the thermal performance of a latent storage energy unit (calci-
um chloride hexahydrate). The latent storage energy unit used polypropylene tubes placed in a calender and
disposed at the center of the greenhouse. The presented results showed that the PCM thermal energy stor-
age unit can maintain the air temperature inside the greenhouse at a value lower than the higher temperature
during a hot day and higher than the night temperature. In addition, for a volume air flow rate of 220 m3·h–1,
the diurnal temperature inside the greenhouse with the PCM thermal energy storage unit was about 49οC for
an ambient temperature equal to 39οC and constant (30oC) during the night (Lazaar et al., 2014). Berroug
et al. (2011) numerically investigated the effect of a PCM thermal energy storage unit on the greenhouse tempera-
ture. The PCM used was calcium chloride hexahydrate placed on the north wall of the greenhouse. The results
showed that the temperatures of plants and air are superior of 6oC to 12oC to the ambient temperature at night in
the winter. In addition, the PCM thermal energy storage unit reduces the interior temperature fluctuations (Berroug
et al., 2011). These results were in agreement with the numerical results of Kumari et al. (2006). The thermal
performance of a phase change thermal storage unit in a greenhouse was investigated experimentally by Benli and
Durmuş (2009). The storage unit consists of a cylindrical tank filled with calcium chloride hexahydrate placed un-
der the ground exterior to the greenhouse. The results showed that the thermal storage unit provoked a difference
between the inside greenhouse temperature and the ambient one varying between 6oC and 9oC.
It is obvious that a work on solar greenhouse dryer with a PCM thermal energy storage unit placed under the
greenhouse dryer has not been studied extensively. So, the aim of this paper is to create a model of the thermal
behavior of a solar greenhouse dryer with a PCM thermal energy storage unit disposed under the interior ground of
the greenhouse. The outline of this paper is the effects of the thermal storage unit on the drying duration of banana
and thermal efficiency of the dryer under Nakorn Pathom (Thailand) climatic conditions.

2. MATHEMATICAL MODEL OF THE GREENHOUSE SOLAR DRYER WITH PCMS STORAGE

2.1 Description

2.1.1 Solar Greenhouse Dryer


The solar greenhouse dryer includes (Fig. 1): (1) a greenhouse in which the products to be dried are placed on trays
and (2) a heat storage unit consisting of a layer of a phase-change material, a layer of concrete, and a metal plate.

FIG. 1: Solar greenhouse dryer with heat storage system

Heat Transfer Research


Solar Greenhouse Dryer with a Phase-Change Material as an Energy Storage Medium 513

The solar greenhouse dryer is made of a polycarbonate film (6 mm of thickness) on a gothic metallic arch. The
dimension of the dryer is 6-m wide, 8-m long, and 3.5-m high. The products to be dried are placed as a thin layer
on four arrays of perforated trays (0.9 × 7 m). These trays are on a metallic plate painted black which acts as an
absorber (αabs = 0.95). The greenhouse is oriented in the north–south direction. The front side of the greenhouse
has two air inlet ports by which the moisture air is sucked out. The ambient air moves into the greenhouse by
forced convection through two inlet ports located on the north side of the greenhouse.
Solar radiation transmitted to the inside of the greenhouse dryer through the cover is partly absorbed by the
products, metal plate, and by the components of the greenhouse dryer. The ambient air is drawn in the greenhouse
through two air inlet ports and heats through the greenhouse by forced convection under the action of solar radi-
ation absorbed by the products and the components of the greenhouse. The drying air flows by forced convection
along the products and absorbs moisture delivered by the products. Solar radiation absorbed by the metal plate is
converted into heat, and this heat will be transferred by conduction to the PCM.

2.1.2 Thermal Storage Unit


The heat storage unit is composed of a layer of PCM of thickness 0.1 m disposed between the metal plate and the
concrete slab (Fig. 2).
The PCM used is paraffin wax because its melting point is equal to 27.7oC that can be reached in the operating
conditions of the solar greenhouse dryer. The latent heat of paraffin wax is equal to 206 kJ·kg–1. The heat solar
flux absorbed at the upper face of the metal plate is transferred by conduction to the PCM and increases its tem-
perature until the melting point. The thermophysical properties of the components of the heat storage system are
given in Table 1.
The heat storage unit involves two processes: charging process and discharging process. In the daytime (charging
process), when the quantity of heat transmitted by conduction through the metal plate to PCMs is sufficient for
reaching the melting point, a portion of PCMs is liquefied and the latent heat of fusion is stored in the PCM liquid
phase. At night or in the absence of sunlight (discharging process), the heat losses between the greenhouse dryer
by convection, radiation and its environment, including radiative transfer with the sky, cause a decrease in the tem-
perature of the components inside the greenhouse dryer. Consequently, the heat transfer by conduction and convec-
tion between PCM's liquid and the air inside the greenhouse causes a decrease in the temperature of the PCM until
the solidification point. The amount of latent heat of solidification is transferred by conduction and convection to
the air inside the greenhouse dryer.

FIG. 2: The form of heat storage system

Volume 49, Issue 6, 2018


514 Aumporn et al.

TABLE 1: Thermophysical properties of the components of the heat storage system


Thermal Conductivity k
Specific Heat Cp (J·kg–1·K–1) Density ρ (kg·m–3)
(W·M–1·K–1)
Metal 26 450 7900
PCM solid 0.18 1800 789
PCM liquid 0.19 2400 750
Concrete 1.4 780 2400

2.2 Modeling

2.2.1 Assumptions
The assumptions in developing the mathematical model for the greenhouse and the thermal storage unit are:
(1) the air flow is unidirectional,
(2) air is transparent to the solar radiation,
(3) the thermal properties of the materials, air, product to be dried, PCM, metal plate, and the concrete are
constant in the temperature range considered,
(4) deformation of the product during drying is neglected,
(5) the sky behaves like a black body,
(6) the temperature and the moisture content of the product are uniform,
(7) the air flow is identical throughout the greenhouse; it means that it remains laminar or turbulent and its
regime does not change,
(8) heat transfer by natural convection in the liquid phase of the PCM is negligible,
(9) PCM behaves ideally, i.e., such phenomena as the property degradation and supercooling are not accounted
for,
(10) PCM is assumed to have a definite melting point (isothermal phase change).

2.2.2 Transfer Equations

• Solar greenhouse dryer


Modeling of the greenhouse dryer uses the time-dependent energy balance and mass balance equations. This
method is based on the analogy between electric transfer and heat transfer. The greenhouse dryer is divided into N
fictive slices perpendicular to the air flow direction (Fig. 3). The amount of heat in a node i is equal to the sum of
the heat fluxes transferred to this node. The energy balance for various components of the greenhouse solar dryer
can be written as

⎛ ∂T JG JJJJG ⎞
mi Cpi ⎜ i + vi grad Ti ⎟ =
⎝ ∂t ⎠
∑hxij Sij (T j − Ti ) + σi , (1)
i

where hxij is the coefficient of heat transfer by the mode x (conduction, convection, radiation) between the media i
and j (W·m–2·K–1); mi is the mass of the component i (kg); Cpi is the specific heat of the component i (J·kg–1·K–1);
σi is the rate of thermal energy absorbed or lost by the component i.
Equation (1) describes different components of the greenhouse solar dryer:

Heat Transfer Research


Solar Greenhouse Dryer with a Phase-Change Material as an Energy Storage Medium 515

a) cover
∂Tc
mc Cpc = hconv, c − amb Ac (Tamb − Tc ) + hr , c − ciel Ac (Tciel − Tc ) + hr , pr − c Ac (T pr − Tc ) + hr , c − f Ac (T f − Tc )
∂t
+ hconv, c − a Ac (Ta − Tc ) + hr , c − se Ac (Tse − Tc ) + ϕα c Ac , (2)
b) drying air
⎡ ∂T ∂Ta ⎞⎤
ma Cpa ⎢ a + Va ⎟⎥ = hconv, a − pr Apr (T pr − Ta ) + hconv, a − f A f (T f − Ta ) + hconv, c − a Ac (Tc − Ta ) , (3)
⎣ ∂t ∂y ⎠⎦

c) product to dry
∂T pr
m pr Cp pr = hconv, a − pr Apr (Ta − T pr ) + hr , pr − c Apr (Tc − T pr )
∂t
(4)
⎛ ∂M ⎞
+ hr , pr − f Apr (T f − T pr ) − Lvap ⎜ ⎟ + ϕτ c α pr Apr ,
⎝ ∂t ⎠
d) metallic plate
∂T f
m f Cp f = hconv, a − f A f (Ta − T f ) + hr , c − f A f (Tc − T f ) + hr , pr − f A f (T pr − T f )
∂t
(5)
+ ϕτ c α f A f + A f hcond (T2 − T f ) ,
jjjjoo
The coefficients of heat transfer by convection and radiation are calculated using correlations given in Kittas
(1985), Halleux (1989), and Duffie and Beckmann (2013).
The drying rate ∂M/∂t is calculated by:
∂M
∂t
(
= Apr β m Cvs (T pr ) − Cv ) (6)

with βm being the mass transfer coefficient determined by Mills (1995):

β m = 0, 332Re 0, 5Sc 0, 33 ; 2000 < Re < 35 · 103; 0.6 < Sc < 2.5 . (7)

FIG. 3: Temperature positions in a slice of greenhouse dryer

Volume 49, Issue 6, 2018


516 Aumporn et al.

In Eq. (6), Cvs(Tpr) and Cv are the saturation water vapor concentration and water vapor concentration of drying
air, respectively. The saturation water vapor concentration is deduced from the sorption isotherm based on the Os-
win modified model (Phoungchandang and Woods, 2000):
C3
1 ⎡ C1 + C2T pr ⎤
= ⎢ ⎥ +1 (8)
HR ⎣ M ⎦

with C1 = 16.68, C2 = – 0.1212, and C3 = 0.9020;


0, 622 Pvs (T pr )
Cvs (T pr ) = , (9)
Pa − 0, 378 Pvs (T pr )

Pv
Pvs (T pr ) = (10)
HR
with Pv being the partial pressure of water vapor in the drying air (Pa).
• Thermal energy storage unit
The solar flux absorbed by the metal plate provokes an increase of its temperature. Then, there is heat propa-
gation by conduction through the metal plate, the PCM, and the concrete slab. Heat transfer in the concrete slab
is described by the fundamental equation of conduction, while heat transfer in the PCM is based on the enthalpy
method modified by Zivkovic and Fujii (2001). At the reference point (0xy) associated with the physical model of
heat storage system, the heat transfer equations can be written by taking account of the above given simplifying
assumptions as follows:
∂Tk ⎛ ∂ 2T ∂ 2Tk ⎞
ρ k C pk = k k ⎜ 2k + ⎟, (11)
∂t ⎜ ∂x ∂y 2 ⎟⎠

where k = 1 for concrete and k = 2 for PCM.


For the PCM, the governing equation is

∂H ⎛ ∂ 2T ∂ 2T2 ⎞
= k 2 ⎜ 22 + ⎟. (12)
∂t ⎜ ∂x ∂y 2 ⎟⎠

In the above equation, the total volumetric enthalpy H is the sum of the sensible and latent heats of PCMs and is
related to the temperature of the PCM as follows:

H = h + ρ 2 L fus f (T2 ) , (13)

where Lf is the latent heat of fusion and f (T2) is the PCM fraction in the liquid state:

⎧ 0 T2 < T f solid

f (T2 ) = ⎨] 0, 1[ T2 = T f fusion . (14)
⎪ 1T > T liquid
⎩ 2 f

The first term on the right-hand side of Eq. (13) represents the volumetric sensible heat h:
Tf
h = ∫Cp 2
dT , (15)
T2

Heat Transfer Research


Solar Greenhouse Dryer with a Phase-Change Material as an Energy Storage Medium 517

T
H (T ) = ∫ ρcdT + ρf ( L fus ) . (16)
Tf

Substituting Eq. (12) into Eq. (13), we obtain

∂h ⎛ ∂ 2T ∂ 2T2 ⎞ ∂f
= k 2 ⎜ 22 + ⎟ − ρ 2 L fus . (17)
∂t ⎜ ∂x ∂y ⎠2 ⎟ ∂t

2.3 Initial and Boundary Conditions

2.3.1 Initial Conditions


∀t < t0; t0 is the time at which the upper surface of metal plate is exposed to solar radiation

Tk ( x, y, t ) = Tinit , (18)

where k = 1 for the concrete, k = 2 for the PCM, k = c for the cover, k = a for the drying air, k = pr for the product,
and k = f for the metal plate.

2.3.2 Boundary Conditions


∀t < t0 ,
∂Tk ⎞
x = 0 and x = L0 : 0 ≤ y ≤ ⎟ = 0 , k = 1, 2 , (19)
∂x ⎠ x = 0, L 0

∂T2 ⎞
y = eb + emcp ; 0 ≤ x ≤ L0 ; hcond (T2 − T f ) = k 2 ⎟ , (20)
∂y ⎠ y = e
b + empc

∂T1 ⎞ ∂T2 ⎞
y = eb ; 0 ≤ x ≤ L0 ; k1 ⎟ = k2 ⎟ , (21)
∂y ⎠ y = e ∂y ⎠ y = e
b b

∂T1 ⎞ ∂Tterre ⎞
y = 0; 0 ≤ x ≤ L0 ; k1 ⎟ = kterre ⎟ . (22)
∂y ⎠ y = 0 ∂y ⎠ y = 0

Two thermal performance parameters are defined: the thermal efficiency of the greenhouse solar dryer and the
efficiency of the heat storage unit during the charging and discharging period.
• The thermal efficiency of the greenhouse solar dryer

Pu
ηg = , (23)
ϕAc

where Pu is the useful power (W), φ is the solar irradiance (W·m–2), and Ac is the area of polycarbonate cover
(m2).

Volume 49, Issue 6, 2018


518 Aumporn et al.

• The storage efficiency of PCM


mmcp f L fus
ηs = , (24)
ϕα f τ c A f

where Af is the area of metal plate (m2), mmcp is the weight of PCM (kg), Lfus is the latent heat of fusion of PCM
(J·kg–1), αf is the absorptivity of metal plate and τc is the transmitivity of polycarbonate cover.
• The retrieval efficiency of PCM

ηd =
(
mmcp , l 1 − f ) L fus
, (25)
mmcp f L fus

where f is the liquid fraction at the end of the charging period.


In order to simulate the thermal behavior of the solar greenhouse dryer with the thermal storage unit during a
day, we supposed that the solar irradiance φ and the ambient temperature evolution during the day are described
by the following expression:
⎛π ⎞
ϕ = ϕ m sin ⎜ (t − HL) ⎟ , (26)
⎝ 24 ⎠

⎛T + Tmin ⎞ ⎛ Tmax − Tmin ⎞ ⎛ 2πt ⎞


Tamb = ⎜ max ⎟−⎜ ⎟ sin ⎜ ⎟, (27)
⎝ 2 ⎠ ⎝ 2 ⎠ ⎝ 24 ⎠
where φm is the maximum solar irradiance (W·m–2); Tmax and Tmin are the maximum and minimum ambient tem-
peratures of the day, respectively, (K); t is the time (h); and HL is the sunrise time (h).

2.4 Numerical Procedure


Equations (1)–(4) are solved using an implicit numerical scheme, the Gauss method associated to an iterative pro-
cedure because the coefficients of heat transfer by convection and radiation depend on the component temperatures
which are unknown. For a time step ∆t, the computations, in a slice of the greenhouse solar dryer, are declared
convergent when the temperatures computed between two successive iterations are inferior to 0.5oC. The drying
air temperature is deduced from the amount of recovered heat by the air through a slice of the greenhouse. This
temperature is supposed to be equal to the inlet one of the next slice.
For the heat storage unit, the dimensionless equations associated with the initial and boundary conditions are
discretized by using the implicit finite difference method. The systems of algebraic equation deduced from these
discretizations are of tridiagonal type. So they were solved using the Thomas algorithm. The house numerical code
was written in Fortran 90.

3. VALIDATION
In order to validate the numerical code developed for heat transfer in PCMs, we applied it to the problem of
Zivkovic and Fujii (2001). These authors performed an experimental and numerical analysis of isothermal phase
change of PCMs within rectangular and cylindrical containers. The experiment is focused on the temperature of
the middle of a parallelepiped tank made of stainless steel (100 × 100 × 20 mm) well insulated on the lateral sides,
filled with a PCM (calcium chloride hexahydrate) (see the thermophysical properties in Table 2). The coefficient
of convective heat transfer hc between the ambient air and the container wall used in our computation is one of
this study and equal to 16 W·m–2·K–1. Figure 4 shows the evolution of the PCM temperature at the center of the
container with time.

Heat Transfer Research


Solar Greenhouse Dryer with a Phase-Change Material as an Energy Storage Medium 519

TABLE 2: Thermophysical properties of CaCl2.6H2O


Melting temperature (oC) — 29.9
–1
Latent heat (kJ·kg ) — 187
Solid 1710
Density (kg·m–3)
Liquid 1530
Solid 1.4
Specific heat (kJ·kg–1·K–1)
Liquid 2.2
Solid 1.09
Thermal conductivity (W·m–1·K–1)
Liquid 0.53

FIG. 4: Evolution of PCMs temperature at the center of the container vs. time

As can be seen from Fig. 4, our results are in good agreement with those of Zivkovic and Fujii (2001). How-
ever, it can be observed that the computed PCMs melting duration is superior to the previous experiment. This
difference seems to be due to the effect of the natural convection within the liquid PCMs that is ignored in the
numerical method.
The modeling of solar greenhouse dryer was validated with the experimental study of Intawee and Janjai (2011).
These authors performed an experimental study for drying banana in a solar greenhouse dryer (7.5 × 20 × 3.7 m)
disposed on the concrete floor. The experiments were started at 8.00 am and continued till 6.00 pm. The drying air
volume flow rate used in this study varied between 400 m3·h–1 and 1200 m3·h–1 and the relative humidity of the
ambient air was in the range from 70% to 80% (Intawee and Janjai, 2011). Figure 5 shows the evolution of the air
temperature inside the solar greenhouse dryer during one day. It can be seen that our results are in good agreement
with those of Intawee and Janjai (2011) with a discrepancy of about 2%. Thus, our numerical code is validated and
can therefore be used to carry out our study.

Volume 49, Issue 6, 2018


520 Aumporn et al.

FIG. 5: Evolution of the air temperature inside the solar greenhouse dryer vs. time

4. RESULTS AND DISCUSSION


The determination of the thickness of the PCM layer used in our modeling has been performed experimentally by
using three containers whose walls except the top are made of plastic (20 × 40 × 10 cm, 20 × 40 × 15 cm, and 20
× 40 × 20 cm). The top of these containers is a metal plate (2 mm thickness). These containers with PCMs were
placed in a solar greenhouse dryer (6 × 8 × 3.5 m) in order to analyze the effect of solar irradiance on their ther-
mal behavior. Three layer thicknesses (10, 15, and 20 cm) of PCM (paraffin wax) have been considered in these
conditions. We observed that the PCM layer of 10-cm thickness is completely melted during the day. In two other
containers, the PCMs are partially melted. Consequently, we retained a PCM layer thickness equal to 10 cm in our
modeling.
Computations were performed for three values of drying air volume flow rate (0.05, 0.1, 0.15 m3·s–1), a relative
humidity of the ambient air (0.9), 3 maximum solar irradiance values (500, 700, and 900 W·m–2), and a minimum
and maximum ambient temperatures equal to 25oC and 30oC. The solar irradiances, air relative humidity, and the
ambient temperature values used in our computation are typical of Nakorn Pathom (Thailand) weather conditions.
The product to be dried is banana (Musa ABB CV. Kluai "Namwa"). The presented results are focused on the ef-
fects of the drying air volume flow rate, the maximum of the solar irradiance on the temperature of the solar green-
house dryer components and that of the heat storage unit. The effects of the drying air volume flow rate on the
solar greenhouse dryer thermal efficiency, the thermal performance of the latent heat thermal storage unit during
the charging and discharging period are also analyzed.

4.1 Influence of the Drying Air Volume Flow Rate


As can be seen in Fig. 6, the moisture content of the product on the first tray decreases quickly during the sun-
shine period (7 am to 5 pm). During the night, the drying process continues with a weak variation of the moisture
content. The next day, as soon as the solar flux acts on the product, the moisture content decreases quickly, as time
increases, to reach the final moisture content (0.25 DB). This figure shows also that the decrease in the moisture
content with time is all the greater as the drying air volume flow rate is higher. In fact, the coefficient of heat and

Heat Transfer Research


Solar Greenhouse Dryer with a Phase-Change Material as an Energy Storage Medium 521

FIG. 6: Evolution of the moisture content of the product vs. time. Influence of drying air volume flow rate

mass transfer between the product and the drying air during the drying process are proportional to the air velocity
and therefore to the drying air volume flow rate. Then, for a given relative humidity of the drying air, the drying
time decreases as the drying air volume flow rate increases. The evolution of the moisture content of the product
along the greenhouse dryer (Fig. 7) shows that at 2 pm the moisture content remains, on the third tray, nearly con-
stant along the greenhouse for the smallest value of the drying air volume flow rate considered in our calculations.
Thus, the drying is not uniform along the greenhouse dryer for the three drying air volume flow rate values. For
the products placed on the trays at the entrance of the dryer, the evaporation is important for the highest value of
the drying air volume flow rate for the relative humidity considered. As the drying air progresses gradually along
the greenhouse dryer, the evaporation decreases because the vapor flux from the products induces an increase of
the moisture content of the drying air and, consequently, reduces the vapor concentration gradient between the
products and the drying air. Diminution of the drying velocity until zero at a distance from the entrance of the
greenhouse dryer will be noted, which is great as the drying air volume flow rate is high. The inflection point on
the curve of the evolution of the moisture content along the greenhouse dryer, for the lowest value of the drying
air volume flow rate, shows a significant increase in the moisture content of the drying air at the entrance of the
greenhouse dryer.
The analysis of Fig. 8 shows that the temperature of the drying air is higher when the drying air volume flow
rate is low. For the drying air volume flow rate equal to 0.05 m3·s–1, the maximum drying air temperature reaches
43oC at noon (Fig. 7). When the inlet conditions for air are fixed, the combined actions of the solar flux absorbed
by the different components of the greenhouse dryer and the convective heat transfer cause an increase of the
drying air temperature and that of the products with time and along the greenhouse dryer. It should be noted that
these temperatures increase with a decrease of the drying air volume flow rate. This result can be explained by the
duration of the air passage through the solar greenhouse dryer, which is even higher when the drying air volume
flow rate is low.

Volume 49, Issue 6, 2018


522 Aumporn et al.

FIG. 7: Evolution of the moisture content of the product along the greenhouse dryer at 14 pm. Influence of drying air volume
flow rate

FIG. 8: Evolution of drying air temperature vs. time. Influence of drying air volume flow rate

The evolution during the day of the temperature of the components of the solar greenhouse dryer and that of the
product is similar to the solar irradiance (Fig. 9). The lowest drying air volume flow rate, we used in this study,
leads to the highest temperature of the solar greenhouse components and that of the drying air and of the product
to be dried. Thus, for the drying air volume flow rate of 0.05 m3·s–1, the temperatures of the banana, the upper face
of the metal plate, and the greenhouse cover reach 40oC, 60oC, and 38oC, respectively.
Figure 10 shows the effect of the drying air volume flow rate on the temperature evolution of the PCM with
time. The maximum temperature of the PCM at the position of the PCM layer (x = 4 m, y = 0.5 m) occurs at 11 am

Heat Transfer Research


Solar Greenhouse Dryer with a Phase-Change Material as an Energy Storage Medium 523

(a)

(b)

(c)

FIG. 9: Evolution of the component temperatures of the greenhouse dryer vs. time. Influence of the drying air volume flow
rate: (a) product, (b) metal plate, and (c) cover

Volume 49, Issue 6, 2018


524 Aumporn et al.

FIG. 10: Evolution of PCM temperature vs. time. Influence of drying air volume flow rate

(TL) as those of the components of the solar greenhouse dryer. For the drying air volume flow rate considered in
this study, the PCM temperature is superior to its melting temperature (27.7oC) during 7 h (8 am to 3 pm). How-
ever, the effect of the drying air volume flow rate on the temperature of the PCM is very weak. This result demon-
strates that the convective heat transfer between the metal plate and the drying air is not predominant comparing to
the solar flux absorbed by the metal plate.

4.2 Thermal Efficiency


The thermal efficiency increases as the drying air volume flow rate increases (Fig. 11) because the amount of heat
recovered by the drying air through the solar greenhouse dryer is much higher. This figure shows also that the
thermal storage unit improves the thermal efficiency of solar greenhouse dryer whatever the drying air volume
flow rate considered in this study. This efficiency varies between 12% and 38% for the greenhouse dryer with the
thermal energy storage unit and between 8% and 28% for the greenhouse dryer without this thermal energy storage
unit. In this case, the lower face of the metal plate is supposed to be adiabatic.
As can be seen from Fig. 12, the thermal efficiency of the solar greenhouse dryer increases as the solar irradi-
ance decreases. As mentioned above, the amount of recovered heat transferred by the drying air through the solar
greenhouse dryer increases, on the one hand, with decrease of solar irradiance and, on the other hand, the drying
air volume flow rate increases.
The thermal efficiency of the storage unit during the charge and discharge periods increases as the solar ir-
radiance decreases (Table 3). These results are corroborated by those about the effect of the drying air volume
flow rate and the solar irradiance on the temperature of the drying air and that of the components of the solar
greenhouse dryer during the day. In fact, for the sunshine duration (7 am to 5 pm), in particular between 7 am to
1 pm, the PCM temperature increases for the low drying air volume flow rate and high solar irradiance. During
this period, the liquid fraction of the PCM increases as the solar irradiance increases. During the discharge period

Heat Transfer Research


Solar Greenhouse Dryer with a Phase-Change Material as an Energy Storage Medium 525

FIG. 11: Evolution of thermal efficiency of the greenhouse dryer with drying air volume flow rate

FIG. 12: Evolution of thermal efficiency of the greenhouse dryer vs. time

(5 pm to 7 am), the drying air temperature decreases until a value inferior to the melting temperature of the PCM.
Consequently, an amount of liquid PCM solidifies which is accompanied by a release of heat transferred by con-
duction through the metal plate and then by convection to the drying air. Thus, the amount of heat released during
the discharge period is greater when the liquid fraction of the PCM at the end of the charge period is low. We
noted that this is the case for the lowest solar irradiance used in this study.

Volume 49, Issue 6, 2018


526 Aumporn et al.

TABLE 3: Thermal efficiency of the PCM storage unit during charging and discharging period: Qv = 0.05 m3·s–1 and
HR = 0.9

Solar Flux Density (W·m–2) Charging Period (7 am–5 pm) (%) Discharging Period (5 pm–7 am) (%)

500 73 27
700 63 8
900 51 3

5. CONCLUSIONS

The thermal behavior of a solar greenhouse dryer with a latent heat storage unit has been studied numerical-
ly under the climatic conditions of Nakorn Pathom (Thailand). The thermal energy storage unit consists of a
PCM layer disposed under the greenhouse. Transfer equations for the solar greenhouse dryer are deduced from
thermal and mass balances performed on its components. In the thermal energy storage unit, transfer equations
are based on the enthalpy method for the PCM layer and classical conduction equation for the concrete layer.
These equations have been solved using an implicit finite method and Gauss algorithm for the equations of
the dryer and Thomas algorithm for those of the thermal storage unit. Results are focused on the effects of the
solar irradiance and drying air volume flow rate on the evolution of the temperature of the components of the
solar greenhouse dryer during a day and that of the thermal storage unit. The thermal performance of the solar
greenhouse dryer and the thermal energy storage unit are also analyzed. The following conclusions can be drawn
for this study:
• The component temperatures of the solar greenhouse dryer (air, product, metal plate, and cover) decrease as
the drying air volume flow rate increases.
• The thermal energy storage unit reduces the drying duration.
• The thermal efficiencies of the solar greenhouse dryer and the thermal storage unit during the charge and
discharge period increase as the solar irradiance decreases.

REFERENCES

Abhat, A., Low Temperature Latent Heat Thermal Energy Storage: Heat Storage Materials, Solar Energy, vol. 30, no. 4,
pp. 313–332, 1983.
Bal, L.M., Satya, S., and Naik, S.N., Solar Dryer with Thermal Energy Storage Systems for Drying Agricultural Food Products:
A Review, Renew. Sustain. Energy Rev., vol. 14, pp. 2298–2314, 2010.
Bal, L.M., Satya, S., Naik, S.N., and Meda, V., Review of Solar Dryers with Latent Heat Storage Systems for Agricultural
Products, Renew. Sustain. Energy Rev., vol. 15, pp. 876–880, 2011.
Bala, B.K., Mondol, M.R.A., Biswas, B.K., Das Chowdury, B.L., and Janjai, S., Solar Drying of Pineapple using Solar Tunnel
Drier, Renew. Energy, vol. 28, pp. 183–190, 2003.
Benli, H. and Durmus, A., Evaluation of Ground-Source Heat Pump Combined Latent Heat Storage System Performance in
Greenhouse Heating, Energy Buildings, vol. 41, pp. 220–228, 2009.
Berroug, F., Lakhal, E.K., El Omari, M., Faraji, M., and El Qarnia, H., Thermal Performance of a Greenhouse with a Phase-
change material North Wall, Energy Buildings, vol. 43, pp. 3027–3035, 2011.
Boulard, T., Razafinjohany, E., Baille, A., Jaffrin, A., and Fabre, B., Performance of a Greenhouse Heating System with a
Phase-Change Material, Agric. Forest Meteorol., vol. 52, pp. 303–318, 1990.

Heat Transfer Research


Solar Greenhouse Dryer with a Phase-Change Material as an Energy Storage Medium 527

Çakmak, G. and Yildiz, C., The Drying Kinetics of Seeded Grape in Solar Dryer with PCM-Based Solar Integrated Collector,
Food Bioproducts Process., vol. 89, pp. 103–108, 2011.
Devahastin, S. and Pitaksuriyarat, S., Use of Latent Heat Storage to Conserve Energy during Drying and Its Effect on Drying
Kinetics of a Food Product, Appl. Thermal Eng., vol. 26, pp. 1705–1713, 2006.
Duffie, J.A. and Beckman, W.A., Solar Engineering of Thermal Processes, 4th ed., New York: John Wiley & Sons, 2013.
Dulawat, M.S. and. Rathore, N.S., Forced Convection Type Solar Tunnel Dryer for Industrial Applications, Agric. Eng. Int.
CIGR J., vol. 14, no. 4, pp. 75–79, 2012.
Esakkimuthu, S., Hassabou, A.H., Palaniappan, C., Spinnler, M., Blumenberg, J., and Velraj, R., Experimental Investigation
on Phase-Change Material Based Thermal Storage for Solar Air Heating Applications, Solar Energy, vol. 88, pp. 144–153,
2013.
Halleux, D. De, Dynamic Model of Heat and Mass Transfer in Greenhouses: Theoretical and Experimental Study, PhD, Gem-
bloux, Belgium, 1989.
Hossain, M.A. and Bala, B.K., Drying of Hot Chilli using Solar Tunnel Drier, Solar Energy, vol. 81, pp. 85–92, 2007.
Intawee, P. and Janjai, S., Performance Evaluation of a Large-Scale Polyethylene Covered Greenhouse Solar Dryer, Int. Energy
J., vol. 12, pp. 39–52, 2011.
Janjai, S., Lamlert, N., Intawee, P., Mahayothee, B., Bala, B.K., Nagle, M., and Muller, J., Experiment and Simulated Per-
formance of a PV-Ventilated Solar Greenhouse Dryer for Drying of Peeled Longan and Banana, Solar Energy, vol. 83,
pp. 1550–1565, 2009.
Kaewkiew, J., Nabnean, S., and Janjai, S., Experimental Investigation of the Performance of a Large-Scale Greenhouse Type
Solar Dryer for Drying Chilli in Thailand, Procedia Engineering, vol. 32, pp. 433–439, 2012.
Kagande, L., Musoni, S., and Madzore, J., Design and Performance Evaluation of Solar Tunnel Dryer for Tomato Fruit Drying
in Zimbabwe, IOSR J. Eng., vol. 2, pp. 01–07, 2012.
Kittas, C., Greenhouse Cover Conductances, Boundary Layer Meteorol., vol. 36, pp. 213–225, 1985.
Kumari, N., Tiwari, G.N., and Sodha, M.S., Effect of Phase-Change Material on Passive Thermal Heating of a Greenhouse, Int.
J. Energy Res., vol. 30, pp. 221–236, 2006.
Kürklü, A., Wheldon, A.E., and Hadley, P., Use of Phase-Change Material (PCM) for Frost Prevention in a Model Greenhouse,
J. Eng. Sci., vol. 3, no. 2, pp. 359–363, 1997.
Lazaar, M., Bouadila, S., Kooli, S., and Farhat, A., Conditioning of the Tunnel Greenhouse in the North of Tunisia using
a Calcium Chloride Hexahydrate Integrated in Polypropylene Heat Exchanger, Appl. Thermal Eng., vol. 68, pp. 62–68,
2014.
Mills, A.F., Basic Heat and Mass Transfer, Massachusetts: Irwin, 1995.
Palled, V., Desai, S.R., Lokesh, L., and Anantachar, M., Performance Evaluation of Solar Tunnel Dryer for Chilly Drying, Kar-
nataka J. Agric. Sci., vol. 25, no. 4, pp. 472–474, 2012.
Patil, R. and Gawande, R., A Review on Solar Tunnel Greenhouse Drying System, Renew. Sustain. Energy Rev., vol. 56,
pp. 196–214, 2016.
Phoungchandang, S. and Woods, J.L., Moisture Diffusion and Desorption Isotherms for Banana, J. Food Sci., vol. 65, no. 4,
pp. 651–657, 2000.
Reyes, A., Mahn, A., and Vasquez, F., Mushrooms Dehydration in a Hybrid-Solar Dryer, using a Phase-Change Material, Ener-
gy Convers. Manage., vol. 83, pp. 241–248, 2014.
Shalaby, S.M. and Bek, M.A., Experimental Investigation of a Novel Indirect Solar Dryer Implementing PCM as Energy Stor-
age Medium, Energy Convers. Manage., vol. 83, pp. 1–8, 2014.
Sharma, A., Tyagi, V.V., Chen, C.R., and Buddhi, D., Review on Thermal Energy Storage with Phase-Change Materials and
Applications, Renew. Sustain. Energy Rev., vol. 13, pp. 318–345, 2009.
Srisittipokakun, N., Kirdsiri, K., and Kaewkhao, J., Solar Drying of Andrographis paniculata using a Parabolic Shaped Solar
Tunnel Dryer, Procedia Engineering, vol. 32, pp. 839–846, 2012.

Volume 49, Issue 6, 2018


528 Aumporn et al.

Zalba, B., Marin, J.M, Cabeza, L.F., and Mehling, H., Review on Thermal Energy Storage with Phase Change Materials: Heat
Transfer Analysis and Applications, Appl. Thermal Eng., vol. 223, pp. 251–283, 2003.
Zivkovic, B. and Fujii, I., An Analysis of Isothermal Phase Change of Phase-Change Material within Rectangular and Cylindri-
cal Containers, Solar Energy, vol. 70, no. 1, pp. 51–61, 2001.

Heat Transfer Research

You might also like