You are on page 1of 27

Thermal efficiency and exergy enhancement of solar air heaters, comparative study

and experimental investigation


Hicham El Ferouali, Ahmed Zoukit, Issam Salhi, Toufiq El Kilali, Said Doubabi, and Naji Abdenouri

Citation: Journal of Renewable and Sustainable Energy 10, 043709 (2018); doi: 10.1063/1.5039306
View online: https://doi.org/10.1063/1.5039306
View Table of Contents: http://aip.scitation.org/toc/rse/10/4
Published by the American Institute of Physics
JOURNAL OF RENEWABLE AND SUSTAINABLE ENERGY 10, 043709 (2018)

Thermal efficiency and exergy enhancement of solar air


heaters, comparative study and experimental investigation
Hicham El Ferouali,1 Ahmed Zoukit,1 Issam Salhi,1 Toufiq El Kilali,2
Said Doubabi,1 and Naji Abdenouri1,a)
1
LSET, Applied Physics Department, Cadi Ayyad University, Marrakesh, Morocco
2
Ste CADVAL 506 Quartier Industriel 40000 Marrakech, Morocco
(Received 7 May 2018; accepted 16 July 2018; published online 24 August 2018)

This paper links the geometric design of most classical configurations of solar air
heaters (SAHs), their thermal performances, and their appropriate application. The
optimization concerns mainly the absorber shape and the airflow direction.
Therefore, single pass and counter flow double pass SAHs with flat plate, finned,
and V-corrugated absorbers and SAH with an external recycle were investigated.
The overall aim is to single out the most suitable configurations for the drying pro-
cess and habitation heating. Modeling of the SAHs was based on solving the energy
balance equations of the SAH’s components. The numerical results showed that the
thermal efficiency was significantly improved by the counter flow double pass mode
by about 17.01% for SAHs with a flat plate absorber and 9.03% for SAHs with a
V-corrugated absorber. Furthermore, the V-corrugated shape of the absorber
increases the thermal and exergy efficiencies compared to the flat plate shape, e.g.,
by 8.66% and 1.27% at m_ ¼ 0.025 kg s1, respectively. In addition, the present
study highlighted that the outlet air temperature and the thermal and exergy efficien-
cies could be enhanced by increasing the number of fins, and the optimal number of
fins was found to be 24 per meter width of the SAH. Experimental studies were car-
ried on a modular SAH. They showed the good capability of the developed model in
predicting the thermal performances of the studied solar collectors. Finally, the
SAHs were assigned to three appropriate uses, namely, aromatic and medicinal plant
drying, agri-food product drying, and habitat heating. Published by AIP Publishing.
https://doi.org/10.1063/1.5039306

NOMENCLATURE
C Specific heat (J kg1 K1)
Dh Hydraulic diameter (m)
f Friction factor
G Solar radiation (W m2)
h Heat transfer coefficient (W m2 K1)
H1 Distance between the glass and the flat plate absorber (m)
H2 Distance between the flat plate absorber and the insulation (m)
H3 Distance between the glass and the upper vertex of the V-corrugated absorber (m)
H4 Height of the V-corrugated absorber (m)
j Node in the flowing air direction
k Thermal conductivity (W m1 K1)
L Length of the SAH (m)
m_ Air mass flow rate (kg s1)

a)
Author to whom correspondence should be addressed: n.abdenouri@uca.ma. Phone: þ212 6 6694 6021. Fax: þ212 5
2443 3170.

1941-7012/2018/10(4)/043709/26/$30.00 10, 043709-1 Published by AIP Publishing.


043709-2 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

N Number of nodes in the flowing air direction


Nu Nusselt number
Pm Pumping power (W)
Pr Prandtl number
R Recycling rate in the case of the SAH with an external recycle (%)
Ra Rayleigh number
Re Reynolds number
t Time (s)
T Temperature (K)
U Bottom heat loss coefficient (W m2 K1)
V Wind velocity (m s1)
Vf Velocity in the air duct (m s1)
W Width of the SAH (m)
x, y, z Spatial co-ordinates (m)
DH Enthalpy change (J kg1)
DS Entropy change (J kg1 K1)

Greek symbols
a Absorptivity
ad Thermal diffusivity of the air (m2 s1)
d Thickness (m)
e Emissivity
t Kinematic viscosity of the air (m2 s1)
geff Effective efficiency of the SAH (%)
gI Thermal efficiency of the SAH (%)
gII Exergy efficiency of the SAH (%)
h Tilt angle of the SAH (rad)
q Density (kg m3)
r Stephen-Boltzmann constant (5.67  108 W m2 K4)
s Transmissivity
(sa) Effective transmittance-absorptance product

Subscripts
a Ambient
(ag) The air gap between the glass and the absorber in the case of the single pass SAH
b Insulation in the bottom
c Convection
f The flowing air in the air duct below the absorber in the case of the single pass
SAH
f1 The flowing air in the air duct above the absorber in the case of the counter flow double
pass SAH
f2 The flowing air in the air duct below the absorber in the case of the counter flow double
pass SAH
fin Fins
g Glass
hc,ik heat transfer by convection from elements i to k (in this paper)
hr,ik heat transfer by radiation from elements i to k (in this paper)
in Inlet air
out Outlet air
p Absorber
r Radiation
043709-3 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

I. INTRODUCTION
For the emergence of solar air heaters (SAHs) on a large scale, the control of temperature
level and efficiency seems to be necessary. SAHs should also support the technology develop-
ment, habitat standards, and small-scale food processing requirements such as drying.
SAHs have the challenge of low thermal capacity of air and low absorber to air heat transfer
coefficient. Mathematical modeling was commonly used to investigate the behavior of the SAHs
versus operational and geometric parameters. Therefore, in previous works,1,2 many mathematical
approaches have been developed for simulating SAHs. Tchinda1 provided a review of the mathe-
matical models that were based on the energy balance analysis. However, it was revealed that
there are still some key deficiencies, and they can only be addressed through continuous enhance-
ment of the existing modeling techniques. In another work, Chamoli et al.2 presented an exten-
sive study of the works carried out on the counter flow and parallel flow double pass SAHs.
Thereby, the recycle ratio is considered an important parameter that affects the performance of
the double pass SAHs, and it is observed that the maximum efficiency is obtained when the
channel depths and air mass flow rate are the same in both the upper and lower ducts.
The geometric shape of the absorber largely influences the thermal performance of the
SAH. Besides, Karim and Hawlader3 showed that the V-corrugated absorber is 10% to 15%
and 5% to 11% more thermally efficient in single pass and counter flow double pass modes
compared to the flat plate absorber, respectively. In other works, Omojaro and Aldabbagh4 con-
firmed these previous results by showing that the counter flow double pass SAH is more ther-
mally efficient than the single pass one by 7% to 19.4%. Furthermore, El-Sebaii et al.5 have
shown that the V-corrugated absorber, compared to the flat plate one, improves the outlet air
temperature by 5% and the thermal efficiency by 11% to 14%. Karim et al.6 investigated theo-
retically and experimentally the counter flow double pass V-corrugated SAH and highlighted
that the solar radiation, the inlet air temperature, and the air flow rate have the most significant
effect on the efficiency of the SAH followed by the air velocity and the SAH’s length, which
has less effect. Wijeysundera et al.7 developed a heat transfer model of the double pass SAH
and compared the results with the single pass one in a wide range of geometrical and operating
conditions. This work emphasized the efficiency rising by 10% to 15% for the double pass
SAH with respect to the single pass one. Choudhury and Garg8 modified the mathematical
model developed in Ref. 7 in order to further investigate the effect of geometrical and operating
parameters. The considered parameters are the length and depth of the flow ducts and the air
mass flow rate. It was reported that the optimum depth of the flow ducts varies significantly
towards lengths and air mass flow rates. Youcef-Ali,9 Moummi et al.,10 and Abene et al.11
have proposed to increase the heat transfer between the absorber and the flowing air by adding
fins to the underside of the absorber. These fins increase the exchange area and reduce the dead
zones.12 Ho et al.13 studied the advantages of performing an external recycle to SAHs and con-
cluded that it increases the forced convection heat transfer. In other works, Mohammadi and
Sabzpooshani14 studied the effect of the external recycle and the insertion of fins and obstacles
on the thermal efficiency of the SAH, and so, they found that the unique use of fins under
external recycle application at high mass flow rates and recycle ratios is an attractive option
that maximizes the thermal efficiency.
Nevertheless, previous investigations were mainly conducted on the thermal performance
of only single structure of SAHs, while comparative studies on the different structures were not
pursued further for clear conclusions. Based only on the literature, optimizing suitable configu-
rations of SAHs for drying and habitat heating seems to be tedious. In fact, previous works
concern only one given type of SAH without integrating, for the same operational conditions,
different absorber shapes, different air flow paths, and different geometries. This study focuses
on gathering most classical types of SAHs and elaborating a dynamic mathematical model
based on energy balance. Then, by performing simulations on the studied SAHs at forced and
natural convection for the same operational conditions, a comparative study was carried out in
terms of the outlet air temperature, the temperature distributions, and the thermal and exergy
efficiencies. The results and trends were checked up by experimental measurements on a novel
043709-4 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

modular solar air heater. In addition, SAHs were allocated to their appropriate use whether dry-
ing or habitat heating.

II. MATERIALS AND METHODS


The experimental apparatus is a modular solar air heater that is used to test all the studied
types, namely, type A: SAH with a flat plate absorber, type B: SAH with a flat plate absorber
and fins, type C: SAH with an external recycle, type D: Single pass SAH with a V-corrugated
absorber, type E: Counter flow double pass SAH with a flat plate absorber, and type F: Counter
flow double pass SAH with a V-corrugated absorber. The modular solar air heater (Fig. 1) was
used to validate the results obtained by the mathematical modeling. It consists of one main
structure composed of the glass and lateral and bottom insulation. Different types of collectors
could be mounted by fixing the appropriate insulated doors and absorbers (Fig. 2). To prevent
against exhaust leaks, all connections are properly fitted with a heat resistant sealing material.
The solar collector with the external recycle was elaborated by using insulated pipes with an
inner diameter of 10 cm and two fans. The fan of 10 cm diameter (Orion 12HBVXC model)
was controlled by the Arduino Uno, allowing the choice of the desired airflow rate. In the case
of the forced convection experiments, the fan sucks the air at the SAH’s outlet. In the case of
the external recycle, two fans were used, the first sucks air from the SAH’s outlet, and the other
one sucks the recycled air at the SAH’s inlet.
PT100 temperature sensors (60.5 K) (ref. TM110, KIMO) were used to measure tempera-
tures at the inlet and the outlet of the SAH. A Kipp&Zonen SMP10 pyranometer (uncertainty
< 1.2%, 12 W/m2 for 1000 W/m2) with a sensitivity of 14.69  103 mV was used to measure
the solar radiation on a 30 inclined surface, and a weather station Vantage Pro2 (ref.
6162CFR) was used to measure the ambient temperature (60.5  C uncertainty) and wind veloc-
ity. The uncertainty of wind velocity is 65% of the measured value. Velocities of drying air
were measured using an anemometer (Kimo model CTV-210-BOS, 0.03 m s1 accuracy).

III. MATHEMATICAL MODELING OF THE STUDIED SAHS


The governing equations were derived by performing the energy balance on each compo-
nent of the SAH as given by the following equation (Ref. 15):

dT
qdC ¼ qin  qout þ qgen ; (1)
dt

FIG. 1. The experimental apparatus: 1. Pyranometer, 2. Modular solar air heater, 3. Anemometer and temperature sensors,
4. Insulated pipes (type C), 5. Power box and fans’ control, and 6. Stable power supply.
043709-5 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

FIG. 2. Synoptic of the modular solar air heater: 1. Glass cover, 2. Absorber, 3. Lower insulated door, 4. Upper insulated
door, 5. Lateral insulation, 6. lower door (types E and F), 7. Upper and lower doors (type C), 8. Upper door (types E and
F), 9. Upper door (types A, B, and D), 10. Lower door (types A, B, and D), 11. V-corrugated absorber, 12. Absorber with
fins, 13. Flat plate absorber, and 14. Insulated pipes (type C).

where q is the density, C is the specific heat, d is the thickness of the layer, dT/dt represents
the variation of temperature (T) vs time (t), qin and qout are the inlet and outlet heat flux per
unit of area, and qgen is the heat generated per unit of area in the analyzed control volume.
The numerical model consists of five main components: the glass cover, an air gap or flow
channel between the glass and the absorber, the absorber, a flow channel below the absorber,
and the bottom insulation.
In the proposed mathematical modeling, the following assumptions were made:
i. Temperatures of the SAHs’ components vary only along the flowing air direction x.16
ii. There is no temperature variation according to the y direction (see Figs. 3–8).17
iii. There is no temperature gradient across the thickness of the glazing, the absorber, and the
insulation.18
iv. Thermo-physical properties of the glass, the absorber, and the insulation are considered
constant.
v. Heat losses from the SAH edges are neglected (According to Duffie and Beckman,19 the
losses from the edges of a large and well insulated SAH could be neglected).
vi. There is no air leakage to or from the SAHs.

A. Energy balance equations of the single pass SAHs


The inlet air gets in the single pass SAHs from one side; it flows between the absorber and
the insulation and gets out from the other side.

1. Energy balance equations of the SAH with a flat plate absorber


The energy balance equations of the SAH with a flat plate absorber (Fig. 3) are given by
the following equations:
For the glass

dTg
qg dg Cg ¼ ag G þ ðhc;ga þ hr;ga ÞðTa  Tg Þ þ hc;gðagÞ ðTðagÞ  Tg Þ þ hr;gp ðTp  Tg Þ: (2)
dt

For the air gap between the glass and the absorber

dTðagÞ
qðagÞ H1 CðagÞ ¼ hc;gðagÞ ðTg  TðagÞ Þ þ hc;pðagÞ ðTp  TðagÞ Þ: (3)
dt
043709-6 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

FIG. 3. Type A: SAH with a flat plate absorber.

FIG. 4. Type B: SAH with a flat plate absorber and fins.

FIG. 5. Type C: SAH with an external recycle.

For the absorber


dTp
qp dp Cp ¼ ap sg G þ hc;pðagÞ ðTðagÞ  Tp Þ þ hr;gp ðTg  Tp Þ þ hc;pf ðTf  Tp Þ þ hr;pb ðTb  Tp Þ:
dt
(4)
043709-7 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

FIG. 6. Type D: Single pass SAH with a V-corrugated absorber.

FIG. 7. Type E: Counter flow double pass SAH with a flat plate absorber.

FIG. 8. Type F: Counter flow double pass SAH with a V-corrugated absorber.

For the air duct

@Tf _ f @Tf
mC
q f H2 C f ¼ hc;pf ðTp  Tf Þ þ hc;bf ðTb  Tf Þ  : (5)
@t W @x
043709-8 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

For the insulation


dTb
qb db Cb ¼ hr;pb ðTp  Tb Þ þ hc;bf ðTf  Tb Þ þ Ub ðTa  Tb Þ: (6)
dt

2. Energy balance equations of the SAH with a flat plate absorber and fins
For the SAH with a flat plate absorber and fins (Fig. 4), the air duct consists of nfin–1 rect-
angular sub-ducts, with nfin being the number of fins.
The convective heat flux from the absorber with fins to the flowing air in the air duct is
calculated according to Bennett and Myers16 through the following equation:

uc;pf ¼ R0 hc;pf Ap;f ðTp  Tf Þ; (7)

with R0 being a dimensionless quantity defined by the following equation:

Afin;f
R0 ¼ 1 þ g ; (8)
Ap;f fin

in which Ap,f is the solid/fluid contact surface between the absorber and the flowing air in the
air duct, and it is equal to ðW  nfin dfin ÞDx.
Afin,f is the solid/fluid contact surface between the fins and the flowing air in the air duct,
and it is equal to 2H2 ðnfin  1ÞDx.
gfin is the efficiency of fins. It is calculated through the following equation as demonstrated
in Refs. 20 and 21:
tanhðMH2 Þ
gfin ¼ : (9)
MH2
By considering dfinL
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffi
2hc;pf ðL þ dfin Þ 2hc;pf
M¼  : (10)
kfin Ldfin kfin dfin

Seeing that Wdfinnfin, the energy balance equations are expressed as follows:
The energy balance equations of the glass, the air gap between the glass and the absorber,
and the insulation are given by (2), (3), and (6), respectively.
For the absorber with fins
dTp
qp ðH2 dfin nfin þ dp WÞCp ¼ ap sg GW þ hc;pðagÞ ðTðagÞ  Tp ÞW þ hr;gp ðTg  Tp ÞW
dt
þhc;pf ðTf  Tp Þð2H2 ðnfin  1Þgfin þ WÞ þ hr;pb ðTb  Tp ÞW: (11)

For the air duct


@Tf @Tf
qf df Cf W ¼ ðW þ 2H2 ðnfin  1Þgfin Þhc;pf ðTp  Tf Þ þ Whc;bf ðTb  Tf Þ  mC
_ f : (12)
@t @x

3. Energy balance equations of the SAH with an external recycle


For the SAH with a flat plate absorber and an external recycle (Fig. 5), the energy balance
equations are the same ones as established for SAH type A except that Eq. (5) is replaced by
Eq. (13)

@Tf _ þ RÞCf @Tf


mð1
qf H2 Cf ¼ hc;pf ðTp  Tf Þ þ hc;bf ðTb  Tf Þ  : (13)
@t W @x
043709-9 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

4. Energy balance equations of the SAH with a V-corrugated absorber


For the SAH with a V-corrugated absorber (Fig. 6), the most effective angle of the V-
corrugated absorber as reported by Karim et al.6 is 60 .
The energy balance equations of the glass and the insulation are given by Eqs. (2) and (6),
respectively.
For the air gap between the glass and the absorber
 
H4 dTðagÞ
qðagÞ H3 þ CðagÞ ¼ hc;gðagÞ ðTg  TðagÞ Þ þ 2hc;pðagÞ ðTp  TðagÞ Þ: (14)
2 dt

For the 60 V-corrugated absorber

dTp
2qp dp Cp ¼ ap sg G þ 2hc;pðagÞ ðTðagÞ  Tp Þ þ hr;gp ðTg  Tp Þ þ 2hc;pf ðTf  Tp Þ þ hr;pb ðTb  Tp Þ:
dt
(15)

For the air duct

H4 @Tf _ f @Tf
mC
qf Cf ¼ 2hc;pf ðTp  Tf Þ þ hc;bf ðTb  Tf Þ  : (16)
2 @t W @x

B. Energy balance equations of the counter flow double pass SAHs


In counter flow double pass SAHs, the inlet air initially flows at the top part of the SAH
above the absorber, and it changes direction once it reaches the end of the SAH and flows
below the absorber to the SAH’s outlet.

1. Energy balance equations of the SAH with a flat plate absorber


The energy balance equations of the counter flow double pass SAH with a flat plate
absorber (Fig. 7) are given by the following equations:
For the glass

dTg
qg dg Cg ¼ ag G þ ðhc;ga þ hr;ga ÞðTa  Tg Þ þ hc;gf 1 ðTf 1  Tg Þ þ hr;gp ðTp  Tg Þ: (17)
dt

For air duct 1 above the absorber

@Tf 1 _ f 1 @Tf 1
mC
qf 1 H1 Cf 1 ¼ hc;pf 1 ðTp  Tf 1 Þ þ hc;gf 1 ðTg  Tf 1 Þ  : (18)
@t W @x

For the absorber

dTp
qp dp Cp ¼ ap sg G þ hc;pf 1 ðTf 1  Tp Þ þ hr;gp ðTg  Tp Þ þ hc;pf 2 ðTf 2  Tp Þ þ hr;pb ðTb  Tp Þ:
dt
(19)

For air duct 2 below the absorber

@Tf 2 _ f 2 @Tf 2
mC
qf 2 H2 Cf 2 ¼ hc;pf 2 ðTp  Tf 2 Þ þ hc;bf 2 ðTb  Tf 2 Þ þ : (20)
@t W @x
For the insulation

dTb
qb db Cb ¼ hr;pb ðTp  Tb Þ þ hc;bf 2 ðTf 2  Tb Þ þ Ub ðTa  Tb Þ: (21)
dt
043709-10 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

2. Energy balance equations of the SAH with a V-corrugated absorber


The energy balance equations of the glass and the insulation (Fig. 8) were given by Eqs.
(17) and (21), respectively,
For air duct 1
 
H4 @Tf 1 _ f 1 @Tf 1
mC
qf 1 H3 þ Cf 1 ¼ 2hc;pf 1 ðTp  Tf 1 Þ þ hc;gf 1 ðTg  Tf 1 Þ  : (22)
2 @t W @x

For the absorber

dTp
2qp dp Cp ¼ ap sg G þ 2hc;pf 1 ðTf 1  Tp Þ þ hr;gp ðTg  Tp Þ þ 2hc;pf 2 ðTf 2  Tp Þ þ hr;pb ðTb  Tp Þ:
dt
(23)

For air duct 2

H4 @Tf 2 _ f 2 @Tf 2
mC
qf 2 Cf 2 ¼ 2hc;pf 2 ðTp  Tf 2 Þ þ hc;bf 2 ðTb  Tf 2 Þ þ : (24)
2 @t W @x

C. Heat transfer coefficients


In the proposed method, the heat transfer coefficients were calculated at every iteration and
for all nodes.

1. Radiation heat transfer coefficients


The radiation heat transfer coefficient between the SAHs’ components was calculated
according to Ref. 19 by the following equation:

rðTi þ Tk ÞðTi2 þ Tk2 Þ


hr;ik ¼ : (25)
1 1
þ 1
ei ek

Between the glass and the sky

reg ðTg4  Ts4 Þ


hr;ga ¼ ; (26)
ðTa  Tg Þ

in which Ts is the sky temperature, estimated by Swinbank22 as

Ts ¼ 0:0552Ta 1:5 : (27)

2. Convective heat transfer coefficients


a. Convection heat transfer coefficient outside the SAHs. The convective heat transfer coeffi-
cient for the air flowing over the outside surface of the glass, according to McAdams,23 is given
by the following equation:

hc;ga ¼ 5:67 þ 3:86V: (28)

b. Natural convection heat transfer coefficients. The natural convection heat transfer coeffi-
cients between the SAHs’ elements are expressed by the following equation:
043709-11 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

hc;ik ¼ Nu:k=H; (29)

in which H is the mean distance between two SAH’s components,


Nu is calculated according to Hollands et al.24 by the following equation:

 þ   " 1 #þ
1708 ðsin 1:8hÞ1:6 1708 Ra cos h 3
Nu ¼ 1 þ 1:44 1  1 þ 1 ; (30)
Ra cos h Ra cos h 5830

with Ra 105.


The Rayleigh number Ra in Eq. (30) is calculated as follows:

gbDTH 3
Ra ¼ ; (31)
tad

with b ¼ 1=T.

c. Forced convection heat transfer coefficients. The forced convection heat transfer coefficients
between the SAHs’ components are expressed below:
The forced convection heat transfer coefficient between a SAH’s component and the air
duct is given by the following equation:

hc;if ¼ Nu:kf =Dh : (32)

First is to calculate the Reynolds number and to determine which Nusselt number equation
should be used [Eq. (33)]

V f Dh
Re ¼ ; (33)
tf

in which Vf is the air velocity in the air duct.


Nu is then calculated as follows:
• For the flat plate absorber
If Re< 2100, according to Mercer et al.,25

0:0606ðRePrDh =LÞ1:2
Nu ¼ 4:9 þ ; (34)
1 þ 0:0909ðRePrDh =LÞ0:7 Pr0:7
in which
qf Cf tf
Pr ¼ : (35)
kf
If Re 2100, according to Kays and Crawford26

Nu ¼ 0:0158Re0:8 : (36)

• For the V-corrugated absorber, according to Hollands and Shewen27


If Re< 2100

Nu ¼ 2:821 þ 0:126ReH4 =L: (37)

If 2800  Re  104

Nu ¼ 1:9  106 Re1:79 þ 225H4 =L: (38)


043709-12 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

If 104<Re<105

Nu ¼ 0:0302Re0:74 þ 0:242Re0:74 H4 =L: (39)

3. Heat losses at the bottom of the SAHs


The bottom heat loss coefficient through the insulation is given by the following equation:

Ub ¼ kb =db : (40)

D. Calculation of the natural convection air mass flow rate


The natural convection air mass flow rate for the SAH types A, B, and D is mainly deter-
mined by two aspects, which are the stack pressure build-up in the air duct and pressure losses
at the inlet, the outlet, and along the air duct. Consequently, the natural convection air mass
flow rate is calculated, according to Al-kayiem et al.,28 by the following equation:
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u2gL sin ðhÞðT
u f ;mean  Ta Þ
m_ ¼ Cd qf Sout u
u  2 ! ; (41)
t Sout
1þ Ta
Sin

Sin and Sout are the inlet and outlet areas of the air duct, respectively. The mean air temperature
in the air duct is calculated, according to Hirunlabh et al.,29 through the following equation:

Tf ;mean ¼ 0:75Tout þ 0:25Tin ¼ 0:75Tout þ 0:25Ta : (42)

The coefficient of discharge Cd, according to Flourentzou et al.,30 is estimated to be equal to 0.6.

E. Thermal performance parameters


The present paper is based on bilateral comparisons, allowing us to cover efficiently all the
studied SAHs. Thermal performances of SAHs are compared based on their outlet air tempera-
tures and their thermal, effective, and exergy efficiencies.

1. Thermal efficiency
The thermal or energy efficiency is defined as the ratio of the energy gain to the solar radi-
ation incident on the SAH [Eq. (43)]31,32

_ f ðTout  Tin Þ
mC
gI ¼ : (43)
LWG

2. Effective efficiency
In order to take into account the fan power consumption, the effective efficiency was intro-
duced by the following equation:

Pm
_ f ðTout  Tin Þ 
mC
geff ¼ C ; (44)
LWG

in which Pm is the pumping power (W), which is the mechanical energy consumption to over-
come the friction. It depends on the pressure drop in the air duct of each collector. C is the
conversion factor accounting the net conversion efficiency from thermal energy to mechanical
energy of pumping power. The value of “C” as recommended by Refs. 33 and 34 is 0.18.
043709-13 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

• For type A
Pm is given by the following equation (Refs. 34 and 35):
2
_
4mfLV f
Pm ¼ : (45)
2Dh

For laminar flow, the friction factor is determined by

24
f ¼ (46)
Re

for turbulent flow Re<105, according to Ref. 36

f ¼ 0:079R0:25
e : (47)

• For type B
The pumping power is determined by the following equation:

Xnfin 1 4m_ i fi LVfi 2 _ i LVf 2


2mf
Pm ¼ ¼ ; (48)
i¼1 2Dhi Dhi

in which fi and Dhi are the friction factor and hydraulic diameter of a sub-duct, respectively. fi is
given by Eqs. (46) and (47).
• For type C
The pumping power is determined by the following formula:

_ þ 1ÞfLVf2 mRf
2mðR _ p LVfp2
Pm ¼ þ ; (49)
Dh Dhp

in which f is given by Eqs. (46) and (47). fp that represents the friction in the recycling pipe is
expressed by
If Rep<2100,
64
fp ¼ : (50)
Rep
If Rep>2100,

fp ¼ ð100Rep Þ0:25 : (51)

• For type D
According to Hollands and Shewen27
 
L 3
3
6m_ f
H4
Pm ¼ : (52)
qðLWÞ2
If Re<2800
H4
13:33
f ¼ þ 0:65 2 : (53)
Re L
If 2800Re104
H4
f ¼ 3:2:10 Re4 0:34
þ 2:94Re 0:19 2 : (54)
L
043709-14 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

If 104Re105
H4
f ¼ 0:0733Re0:25 þ 0:51 2 : (55)
L
• For type E:
The pumping power is determined by the following equation:

_ 1 LVf 1 2 2mf
2mf _ 2 LVf 2 2
Pm ¼ þ ; (56)
Dh1 Dh2

in which f1 and f2 are the friction factors in the upper and lower ducts, respectively. They are
determined by Eqs. (46) and (47).
• For type F
The pumping power is determined by the following equation:
0 13
3 L
6m_ f1 B C 
@
H3 þ
H4 A 3 L 3
6m_ f2
2 H4
Pm ¼ þ : (57)
q1 ðLWÞ2 q2 ðLWÞ2

If Re<2800

H3 þ H4
13:33 2
f1 ¼ þ 0:65 : (58)
Re1 L

If 2800Re104

H4
H3 þ
f1 ¼ 3:2  104 Re1 0:34 þ 2:94Re1 0:19 2 : (59)
L

If 104  Re  105

H4
H3 þ
f1 ¼ 0:0733Re1 0:25
þ 0:51 2 : (60)
L

f2 is given by Eqs. (53)–(55).

3. Exergy efficiency
The exergy efficiency is a paramount criterion in the optimization and performance evaluation
of energy systems.
The exergy efficiency of a SAH was evaluated by taking into account the fan power con-
sumption as37

Ta
_
mðDH  Ta DSÞ  Pm
Tf ;mean
gII ¼   ; (61)
Ta
1 ðsaÞLWG
Tso

in which
043709-15 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

• Tso¼ 5600 K is the solar temperature.


• (sa) is the effective product transmittance-absorptance that depends mainly on the transmissiv-
ity of the transparent glass and on the absorptivity of the absorber. It is equal to the optical effi-
ciency, and it is estimated by the following equation:38

ðsaÞ  1:2sg ap : (62)

• The enthalpy change DH and entropy change DS of the air in the SAH are expressed by the fol-
lowing equations:38–41

DH ¼ Cf ðTout  Tin Þ; (63)

Tout Pout
DS ¼ Cf ln  Rs ln ; (64)
Tin Pin

where Pin(Pa) is the fluid pressure at the SAH’s inlet and Pout(Pa) is the fluid pressure at its
outlet. Rs ¼ 287.058 J kg1 K1 is the universal gas constant.
Pout qout Tout
¼ : (65)
Pin qin Tin

F. Numerical solution of the mathematical model


In numerical solution of the energy balance system, systems of differential equations were
solved using the implicit finite difference method. In the proposed method, all temperatures
should check the error criterion to stop the iteration process. The program’s input includes design
and operational parameters. Tables I and II show the thermo-physical and operational parameters
employed in the numerical calculations, respectively. In the proposed numerical method, the
thermo-physical properties of the air are temperature-dependent. Based on their values at some
temperatures; the cubic spline interpolation was used to determine them at every temperature.
The geometric dimensions employed in the numerical calculations are given in Table III. Forced
convection simulations were performed at an air mass flow rate of 0.025 kg s1 which is the rec-
ommended value for solar air heaters38,42 in order to enhance their thermal performances.
The proposed numerical solution is able to predict the instantaneous temperatures of the
SAH’s components at all nodes. From the temperature evolutions versus time, it is noticed that
the collectors, under constant operating conditions, behave as first order systems.
The computed temperatures versus time of type A’s components, at the nodes j ¼ N/2 and
j ¼ N, are presented in Figs. 9(a) and 9(b), respectively. The curves’ shape at the node j ¼ N/2
is similar to the one at j ¼ N, except that the achieved range of temperature is higher at the
node j ¼ N. This trend is verified for all the other SAH types.
The temperature curves versus time of the SAHs, types B, C, D, E, and F, at the node N/2
have a similar trend to the ones given in Fig. 9. For these types, the biggest 95% response time

TABLE I. Thermo-physical parameters employed in the numerical calculations.43,44

Parameter Value Parameter value

sg 0.94 eb 0.6
ag 0.04 Cg (J kg1 K1) 720
eg 0.1 Cp(J kg1 K1) 385
ap 0.96 Cb (J kg1 K1) 1400
ep 0.12 qg(kg m3) 2500
1 1
kb (W m K ) 0.025 qp(kg m3) 8940
kfin (W m1 K1) 390 qb (kg m3) 32
043709-16 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

TABLE II. Operational parameters employed in the numerical calculations.

Parameter Value Parameter Value

G (W m2) 700 V (m s1) 5


1
Ta ( C) 25 m_ (kg s )(in forced convection) 0.025
h (deg) 30

TABLE III. Geometric dimensions employed in the numerical calculations.

Parameter Value (m) Parameter Value (m) Parameter Value (m)

L 2 H3 0.025 db 0.05
W 1 H4 0.05 dfin 0.001
H1 0.05 dg 0.003
H2 0.025 dp 0.0005

in which temperatures reaches 95% of the threshold value is the insulation’s one. It is equal to
24, 27, 22, 23, 30, and 28 min for the SAH types A, B, C, D, E, and F, respectively.

IV. RESULTS AND DISCUSSION


A. Comparison between the SAHs with flat plate and V-corrugated absorbers in single
and counter flow double pass modes
1. Comparison between the SAHs with flat plate and V-corrugated absorbers in single
pass modes
It is shown that the absorber temperature of type A is widely higher compared to the one of
type D in both forced and natural convection (Fig. 10). On the other hand, the flowing air tem-
perature is higher for type D compared to type A (Fig. 11). Hence, the corrugated shape of the
type D’s absorber enhances the heat transfer from the absorber to the flowing air in the air duct.
Effective efficiency and thermal efficiency curves are superimposed [Fig. 12(a)]. The dif-
ference between the two efficiencies is only 0.01% for both types A and D. Hence, the pressure
drop is negligible in the air ducts. According to Figs. 12(a) and 12(b), the SAH with a V-
corrugated absorber (type D) is more energy efficient than the one with a flat plate absorber
(type A) by about 8.66% in forced convection and 5.73% in natural convection. It is more
exergy efficient by about 1.27% in forced convection and 0.83% in the natural convection. This
result shows that the use of the V-corrugated absorber improves significantly the thermal per-
formances of the SAH, which is in agreement with previous results.3,5,45 In addition to the fact
that this improvement of the efficiencies is due to the enhancement of the convective heat

FIG. 9. Temperature of different components of SAH type A, (a) at node j ¼ N/2 and (b) at node j ¼ N.
043709-17 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

FIG. 10. Comparison between the absorber temperatures of types A and D in forced and natural convection.

FIG. 11. Comparison between the flowing air temperatures of types A and D in forced and natural convection.

transfer coefficient between the V-corrugated absorber and the flowing air in the air duct, the
use of the V-corrugated absorber increases the contact surface between it and the flowing air
and creates turbulence resulting in larger heat gains for the fluid.46

2. Comparison between counter flow double pass and single pass SAHs
For types E and F, the absorber temperature decreases with the increase in x (Figs. 13 and
14). Contrariwise, the absorber temperature of the SAH types A and D increases with the increase
in x. Besides, the flowing air in types E and F, in contrast to types A and D (Figs. 13 and 14),
undergoes a pre-heating in air duct 1, then, it changes the direction once it reaches the end of the
SAH, and its temperature continues to increase till the SAH’s outlet. Therefore, type E benefits
from more contact time with the absorber compared to type A. The same results were obtained in
the comparison between the SAH types D and F. As a result, type E is higher than type A, and
type F is higher than type D in terms of the outlet air temperature. Moreover, it was found that
counter flow double pass SAHs are more energy efficient than the single pass ones by about
17.01% for SAHs with a flat plate absorber and 9.03% for SAHs with a V-corrugated absorber.
Almost, the same rates with a difference of 0.01% were found in terms of the effective efficiency.
Thus, counter flow double pass SAHs, as also reported by Omojaro and Aldabbagh,4 increase the
043709-18 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

FIG. 12. (a) The evolution of thermal, effective, and exergy efficiencies of types A and D versus time in forced convection,
(b) The evolution of thermal and exergy efficiencies of types A and D versus time in Natural convection.

FIG. 13. Comparison between types A and E in terms of the absorber and the flowing air temperature distributions.

FIG. 14. Comparison between types D and F in terms of the absorber and the flowing air temperature distributions.
043709-19 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

FIG. 15. Comparison between the temperature distributions of type B and type D: (a) Forced convection and (b) Natural
convection.

thermal efficiency. Furthermore, they are more exergy efficient than the single pass ones by about
2.63% and 1.48% for the SAHs with flat plate and with V-corrugated absorbers, respectively.

B. Comparison between the SAH with fins and the SAH with a V-corrugated absorber
In order to compare between the SAH with fins (type B) and the SAH with a V-corrugated
absorber (type D), the contact area between the absorber and the flowing air in the air duct
must be equal for the two configurations. The number of fins that satisfies this condition is 21.
The absorber temperature is higher for type B compared to type D in both forced and natu-
ral convection (Fig. 15). Moreover, given that Tf values are equal in forced convection and
almost equal in natural convection and that T(ag) is higher for type D, type D enables more heat
transfer from the absorber to the air gap between the absorber and the glass. The thermal and
exergy efficiencies are equal for the two configurations in forced convection with differences of
0.22% and 0.03%, respectively. In natural convection; they are slightly superior for type B by
only about 2.71% and 0.41%, respectively.

C. Study of the effect of fins


The thermal and exergy efficiencies are computed in the steady state mode for different
numbers of fins. It is noticed that the thermal and exergy efficiencies (Fig. 16) of the SAH with
2 fins are slightly higher than those of the SAH without fins (type A), and the addition of fins
increases these factors. The SAH with fins becomes more efficient than the one with the V-
corrugated absorber (type D) when the number of fins exceeds 20 fins in the forced convection

FIG. 16. Thermal and exergy efficiencies of type B versus the number of fins: (a) Forced convection and (b) Natural
convection.
043709-20 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

and 12 fins in the natural convection. In addition, the thermal efficiency and the exergy effi-
ciency of type B become stable when the number of fins exceeds 24 fins in both cases forced
and natural convection. Hence, the optimal number of fins for type B is 24 fins.

D. Study of the effect of the external recycle


The SAH with an external recycle was operated at three different flow rates: 0.015, 0.025, and
0.035 kg s1. The thermal and exergy efficiencies increase by increasing the recycle rate (Fig. 17).
But this increase is very weak for the three different flow rates indeed with a recycling rate of
80%, the maximum increase in the thermal efficiency and the exergy efficiency is 3.85% and
0.54%, respectively, occurring at the air flow rate of m¼ _ 0.025 kg s1. In fact, the low temperature
difference between the SAH’s inlet and outlet in the case of the external recycle does not allow the
air to recover enough energy that will meaningfully increase temperature. In addition, the thermal
losses of the inlet air that is already hot increases throughout the SAH.

V. EXPERIMENTAL VALIDATION AND DISCUSSION


Eleven experiments were performed in forced and natural convection during January and
February 2018 in Marrakech (Morocco). In the case of the forced convection experiments, an
air flow rate of 0.025 kg s1  80 m3 h1 was used, and three recycling rates were tested for
type C (20%, 50%, and 80%). The experiments were started at the solar time 12:13 6 5 min
(local time 11:30 am) and finished at 14:13 6 5 min (13:30 pm). During this period, solar radia-
tion and ambient temperature do not vary significantly. Before 11:30 am, the modular SAH was
kept in shadow. Temperature, solar radiation, and wind velocity were recorded every 5 min.
The properties of the used materials in the designed apparatus are for the glass: dg ¼ 6 mm,
sg ¼ 0.9, eg ¼ 0.85, qg ¼ 2500 kg m3, Cg ¼ 720 J kg1 K1, and ag ¼ 0.04. The absorber is
made from aluminum covered with matt black glycerophtalic lacquer (L ¼ 2 m, W ¼ 1 m,
dp ¼ 0.5 mm, ap ¼ 0.95, ep ¼ 0.8, qp ¼ 2688.9 kg m3, and Cp ¼ 897 J kg1 K1). For the
absorber with fins, the number of fins is Nfin ¼ 24, and they are made from aluminum
(dfin ¼ 1 mm and kfin ¼ 237 W m1 K1). The insulation in the bottom, the edges of the SAHs,
and the insulated doors were made from a steel sheet metal and were isolated by mineral fiber
of thickness 40 mm (db ¼ 4 cm, qb ¼ 30 kg m3, kb ¼ 0.035 W m1 K1, Cb ¼ 1000 J kg1 K1,
and eb ¼ 0.6). These thermo-physical properties related to the designed SAH were used in the
numerical calculations for the experimental validation part.
Figure 18 presents the solar radiation and ambient temperature during the eleven day
experiments. All experimental data are gathered in one figure. Solar radiation forms a narrow
cloud point with a mean value of 837 6 36.7 W m2. In addition, all the recorded values for

FIG. 17. Effect of the external recycle on the thermal and exergy efficiencies.
043709-21 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

FIG. 18. Solar radiation and ambient temperature for the eleven day experiments.

ambient temperature range between almost 11 and 20  C. Wind velocity was almost constant.
Its mean value is almost 3 m s1. This value is used in the numerical calculation of this valida-
tion part.
Figure 19 shows the experimental and theoretical data of the outlet air temperature in the
forced convection mode. Its mean relative error, calculated in ( C), ranges between 2.75% and
5.13%. In terms of the temperature difference (Tout – Tin), the mean relative error ranges
between 3.96% and 5.46%.
Figure 20 shows the experimental and theoretical outlet air temperatures and the air flow
rate in the natural convection mode; the mean relative error ranges between 3.02% and 4.62%
for Tout, between 4.50% and 6.97% for (ToutTin), and between 5.06% and 5.95% for the air
flow rate. Furthermore, Figs. 21 and 22 present the experimental and theoretical thermal effi-
ciency in forced and natural convection, respectively, and its mean relative error ranges
between 2.01% and 5.17% in forced convection and between 4.63% and 6.79% in natural
convection.
The studied SAHs’ cost varies according to their specificity. In other words, the 60 V-
corrugated shape of the absorber (types D and F) requires twice the surface area of the flat plate

FIG. 19. The experimental and theoretical outlet air temperature in forced convection.
043709-22 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

FIG. 20. The experimental and theoretical outlet air temperature and air flow rate in natural convection.

absorber (types A and E), as well as a specific bending. In addition, the geometrical shape of
the finned absorber (type B) requires additional welding and bending costs. The SAH with an
external recycle needs more investment in terms of controlled fans and insulated pipes. The
additive cost for the realization of the different configurations compared to type A was calcu-
lated. They do not exceed 20% of type A’s cost. Therefore, the cost aspect could be neglected
and comparisons between the different types could be focused only on their thermal performan-
ces and to their appropriate application.
The experimental effective efficiency and exergy efficiencies were determined using Eqs.
(44) and (61), respectively, and by assuming that the thermal power Pm/C equals the consumed
power of the fan Pfan.
Tables IV and V confirm that counter flow double pass SAHs improve significantly the
studied thermal performances. In fact, they are more thermal, effective, and exergy efficient
than the single pass ones by 16.89 6 2.94%, 16.91 6 3.11%, and 1.75 6 0.17% for the flat plate
absorber and around 9.64 6 3.09%, 9.64 6 3.13%, and 1.06 6 0.20% for the V- corrugated
absorber, respectively, which is in accordance with the theoretical results. In addition, type D is
higher than type A in terms of thermal efficiency and is about 8.31 6 2.75% in forced convec-
tion and 5.59 6 1.19% in natural convection. It is more exergy efficient by about 0.82 6 0.13%
in forced convection and 0.67 6 0.13% in the natural convection. Moreover, by comparing
types D and B, it is found that the thermal and exergy efficiencies are almost equal for both

FIG. 21. The experimental and theoretical thermal efficiency in forced convection.
043709-23 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

FIG. 22. The experimental and theoretical thermal efficiency in natural convection.

types in the forced convection mode, and they are slightly higher for type B in the natural con-
vection mode by 3.57 6 1.23% and 0.38 6 0.18%, respectively. Finally, the experimental results
confirm that the outlet air temperature, the thermal, effective, and exergy efficiencies increase
by increasing the recycling rate, but this enhancement is not significant.
According to Tables IV and V, for climatic conditions in which solar radiation is about
837 6 36.7 W m2 and for moderate ambient temperature ranging between almost 11 and
20  C, the solar collectors could be assigned to three appropriate applications. Types A and C
could be applied for drying aromatic and medicinal plants requiring moderate temperatures
lower than 50  C in order to protect sensitive active ingredients.47 The SAH types B and D in
forced convection and type D in natural convection are more suitable for drying agri-food prod-
ucts since their recommended drying temperatures are between 50 and 60  C. Besides, the opti-
mal SAHs for habitat heating are those that record the highest thermal efficiency (Figs. 21 and
22) and effective and exergy efficiencies since the heating is especially used at insufficient solar
radiation and low ambient temperature. Hence, the SAHs with types E and F in forced convec-
tion and type B in natural convection are more appropriate for habitat heating because they
recorded the highest gI (%), geff (%), and gII (%).

VI. CONCLUSIONS
In the present paper, six types of single pass and counter flow double pass SAHs with flat
plate, finned, and V-corrugated absorbers and with an external recycle were investigated by
elaborating a dynamic mathematical model for each type based on energy balance. The effec-
tive efficiency of the studied collectors is almost equal to the thermal efficiency, showing that
the pumping work of the fan is negligible. By performing bilateral comparisons between the
studied SAHs, it was found that counter flow double pass SAHs improve significantly the stud-
ied thermal performances. In fact, they are more thermally and exergy efficient than the single
pass ones by 17.01% and 2.63% for the flat plate absorber and around 9.03% and 1.48% for
the V-corrugated absorber, respectively. In addition, the V-corrugated shape of the absorber
increases the thermal and exergy efficiencies compared to the flat plate shape, e.g., by 8.66%
and 1.27% at m_ ¼ 0.025 kg s1, respectively. Furthermore, the studied thermal performances
increase by increasing the number of fins, and 24 fins per meter width of the SAH is the opti-
mal number for the considered dimensions in this paper. Moreover, the external recycle does
not lead to a significant gain in the thermal and exergy efficiencies.
Experimental studies were carried on a modular solar air heater in order to test all the stud-
ied configurations, validate the mathematical models, and confirm the obtained results by the
mathematical modeling tool. The mean relative error ranges between 2.75% and 4.62% for the
outlet air temperature, and it ranges between 5.06% and 5.95% for the air flow rate, which
demonstrates the good capability of the developed model for predicting the thermal
043709-24
El Ferouali et al.
TABLE IV. Experimental values for G(w/m2), Ta ( C), Tout ( C), gI (%),geff (%), and gII (%) in forced convection.

SAHs in forced convection (m_ ¼ 0.025 6 0.001 kg s1)

Parameter Type A Type B Type C R ¼ 20% Type C R ¼ 50% Type C R ¼ 80% Type D Type E Type F
2
G (W m ) 813.3 6 9.7 849.5 6 10.1 860.9 6 10..3 859.3 6 10.3 832.8 6 9.9 847.7 6 10.1 823.2 6 9.8 837.2 6 10.01
Ta ( C) 11.86 6 0.50 20.72 6 0.50 12.28 6 0.50 11.98 6 0.50 13.47 6 0.50 18.91 6 0.50 16.22 6 0.50 18.24 6 0.5
Tout ( C) 40.01 6 0.50 55.83 6 0.50 44.08 6 0.50 46.81 6 0.50 48 6 0.50 53.8 6 0.50 55.7 6 0.50 59.1 6 0.50

J. Renewable Sustainable Energy 10, 043709 (2018)


gI (%) 43.53 6 1.84 52.07 6 2.05 46.50 6 1.89 51.04 6 2.02 52.20 6 2.07 51.84 6 2.05 60.42 6 2.30 61.48 6 2.32
geff (%) 43.50 6 2.04 52.06 6 2.12 46.49 6 2.09 51.03 6 2.04 50.18 6 2.08 51.83 6 2.09 60.41 6 2.35 61.47 6 2.34
gII (%) 2.02 6 0.08 2.86 6 0.09 2.39 6 0.11 2.84 6 0.12 2.84 6 0.12 2.84 6 0.11 3.77 6 0.15 3.90 6 0.17
043709-25 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

TABLE V. Experimental values for G (W/m2), Ta ( C), Tout ( C), gI (%), gII (%), and m_ in natural convection.

SAHs in natural convection

Parameter Type A Type B Type D


2
G(W m ) 795.1 6 9.5 838.4 6 10.1 849.3 6 10.2
Ta ( C) 14.76 6 0.50 18.38 6 0.50 19.08 6 0.50
Tout ( C) 49.19 6 0.50 61.16 6 0.50 59.80 6 0.50
_
m(kg s1) 0.0163 6 0.0008 0.0174 6 0.0008 0.0170 6 0.0008
gI (%) 35.57 6 0.82 44.73 6 0.88 41.16 6 0.87
gII (%) 1.97 6 0.07 3.02 6 0.14 2.64 6 0.12

performance of the studied solar collectors. By comparing the experimental thermal, effective,
and exergy efficiencies of the different studied types, all theoretical comparison trends were
confirmed. Besides, the studied SAHs could be assigned to three appropriate uses for climatic
conditions in which solar radiation is about 837 6 36.7 W m2 and for moderate ambient tem-
perature ranging between almost 11 and 20  C. Types A and C could be applied for drying aro-
matic and medicinal plants. The SAH types B and D in forced convection and type D in natural
convection are more suitable for drying agri-food products. Finally, types E and F in forced
convection and type B in natural convection are appropriate for habitat heating.

ACKNOWLEDGMENTS
This work was supported by the research institute IRESEN as part of the project SSH, and all
the authors are grateful to the IRESEN institute for its cooperation.
1
R. Tchinda, Renewable Sustainable Energy Rev. 13, 1734 (2009).
2
S. Chamoli, R. Chauhan, N. S. Thakur, and J. S. Saini, Renewable Sustainable Energy Rev. 16, 481 (2012).
3
M. A. Karim and M. N. A. Hawlader, Energy 31, 452 (2006).
4
A. P. Omojaro and L. B. Y. Aldabbagh, Appl. Energy 87, 3759 (2010).
5
A. A. El-Sebaii, S. Aboul-Enein, M. R. I. Ramadan, S. M. Shalaby, and B. M. Moharram, Energy 36, 1076 (2011).
6
M. A. Karim, E. Perez, and Z. M. Amin, Renewable Energy 67, 192 (2014).
7
N. E. Wijeysundera, L. L. Ah, and L. E. Tjioe, Sol. Energy 28, 363 (1982).
8
C. Choudhury and H. P. Garg, Renewable Energy 1, 595 (1991).
9
S. Youcef-Ali, Renewable Energy 30, 271 (2005).
10
N. Moummi, S. Youcef-Ali, A. Moummi, and J. Y. Desmons, Renewable Energy 29, 2053 (2004).
11
A. Abene, V. Dubois, M. Le Ray, and A. Ouagued, J. Food Eng. 65, 15 (2004).
12
S. Rai, P. Chand, and S. P. Sharma, Renewable Energy 125, 39 (2018).
13
C. D. Ho, H. M. Yeh, and T. C. Chen, Int. Commun. Heat Mass Transfer 38, 49 (2011).
14
K. Mohammadi and M. Sabzpooshani, Energy Convers. Manage. 88, 239 (2014).
15
R. Perez-Espinosa and O. Garcıa-Valladares, J. Renewable Sustainable Energy 10, 013705 (2018).
16
C. D. Ho, H. M. Yeh, and R. C. Wang, Energy 30(15), 2796 (2005).
17
K. K. Matrawy, Sol. Energy 63, 191 (1998).
18
M. R. I. Ramadan, A. A. El-Sebaii, S. Aboul-Enein, and E. El-Bialy, Energy 32, 1524 (2007).
19
J. A. Duffie and W. A. Beckman, Solar Engineering of Thermal Processes, 2nd ed. (John Wiley & Sons, Inc., 2013).
20
C. O. Bennett and J. E. Myers, Momentum, Heat and Mass Transfer (McGraw-Hil, New York, 1962).
21
H. M. Yeh, J. Taiwan Inst. Chem. Eng. 43, 235 (2012).
22
W. C. Swinbank, Q. J. R. Meteorol. Soc. 89, 339 (1963).
23
W. H. McAdams, Heat Transmission, 3rd ed. (McGraw-Hill, New York, 1954).
24
K. G. T. Hollands, T. E. Unny, G. D. Raithby, and L. Konicek, J. Heat Transfer 98, 189 (1976).
25
W. E. Mercer, W. M. Pearce, and J. E. Hitchcock, J. Heat Transfer 89, 251 (1967).
26
W. M. Kays and M. E. Crawford, Convective Heat and Mass Transfer, 2nd ed. (McGraw-Hill, New York, 1980).
27
K. G. T. Hollands and E. C. Shewen, J. Sol. Energy Eng. 103, 323 (1981).
28
H. H. Al-kayiem, A. K. Hussein, and T. S. Peow, Int. J. Mech. Aerosp., Ind. Mechatronics Eng. 7(3), 452–457 (2013).
29
J. Hirunlabh, W. Kongduang, P. Namprakai, and J. Khedari, Renewable Energy 18, 109 (1999).
30
F. Flourentzou, J. Van der Maas, and C.-A. Roulet, Energy Build. 27, 283 (1998).
31
H. F. Oztop, F. Bayrak, and A. Hepbasli, Renewable Sustainable Energy Rev. 21, 59 (2013).
32
M. Sekar, M. Sakthivel, S. Satheesh Kumar, and C. Ramesh, J. Renewable Sustainable. Energy 4, 042901 (2012).
33
A. Cortes and R. Piacentini, Appl. Energy 36, 253 (1990).
34
A. Kumar, R. P. Saini, and J. S. Saini, Renewable Sustainable Energy Rev. 37, 100 (2014).
35
A. A. Hegazy, Energy Convers. Manage. 41, 1361 (2000).
36
C. D. Ho, H. M. Yeh, T. W. Cheng, T. C. Chen, and R. C. Wang, Appl. Energy 86, 1470 (2009).
043709-26 El Ferouali et al. J. Renewable Sustainable Energy 10, 043709 (2018)

37
M. Hedayatizadeh, F. Sarhaddi, A. Safavinejad, F. Ranjbar, and H. Chaji, Energy 94, 799 (2016).
38
H. Esen, Build. Environ. 43, 1046 (2008).
39
Y. A. Çengel and M. A. Boles, Thermodynamics: An Engineering Approach, 5th ed. (McGraw-Hill, New York, 2006).
40
S. Karsli, Renewable Energy 32, 1645 (2007).
41 _
I. Kurtbas and A. Durmuş, Renewable Energy 29, 1489 (2004).
42
F. Bayrak and H. F. Oztop, J. Therm. Sci. Technol. 35, 11 (2015).
43
H. Tabord, Solar Energy Conversion (Elsevier, 1979), pp. 253–286.
44
J. F. Saccadura, Initiation Aux Transfert Thermique (Technique et documentation, Paris, 1978).
45
T. Liu, W. Lin, W. Gao, C. Luo, M. Li, Q. Zheng, and C. Xia, Int. J. Green Energy 4, 601 (2007).
46
S. Li, H. Wang, X. Meng, and X. Wei, Appl. Therm. Eng. J. 114, 639 (2017).
47
J. M€uller and A. Heindl, Drying of Medicinal Plants (Frontis, 2006).

You might also like