Você está na página 1de 14

An experimental investigation of an inclined passive wall solar

chimney for natural ventilation


Rakesh Khanal

, Chengwang Lei
School of Civil Engineering, The University of Sydney, Sydney, NSW 2006, Australia
Received 13 November 2013; received in revised form 7 March 2014; accepted 19 May 2014
Available online 28 June 2014
Communicated by: Associate Editor S.A. Sherif
Abstract
Ongoing investigations into solar chimney development have resulted in constantly evolving new designs. In this study, experiments
are carried out with an inclined passive wall solar chimney (IPWSC) model with a uniform heat ux on the active (absorptive) wall. The
eectiveness of this design has been examined for the heat ux range of 100 W/m
2
500 W/m
2
with a xed base air gap width of 0.1 m and
inclination angles of the passive wall in the range of 06 degrees. The experimental results show that the inclination angle of the passive
wall has no signicant eect on the temperature distribution across the air gap width and along the chimney height. On the other hand,
the averaged air ow velocity across the air gap width is strongly aected by the inclination angle. The experimental results also show
that the IPWSC with 0.7 m absorber height and 0.1 m air gap width at an inclination angle of 6 and input heat ux of 500 W/m
2
can
produce sucient ventilation for a 27 m
3
room based on ASHREA standards. Further, the present experimental results show that the
IPWSC design can signicantly improve the ventilation performance of a solar chimney in comparison to the conventional chimney
design with vertical passive wall conguration. The experimental results are supported by ow visualization experiments and are consis-
tent with scaling predictions.
2014 Elsevier Ltd. All rights reserved.
Keywords: Natural ventilation; Solar chimney; Scale analysis; Flow visualization
1. Introduction
Natural ventilation in building has attracted a growing
research interest, especially in the last two decades because
of increasing environmental concern regarding greenhouse
gas emission and the need for an ecient and eective
passive ventilation system as part of the green building
architecture. In this regard, solar chimney oers a promis-
ing solution. However, a recent review (Khanal and Lei,
2011a) on the subject has revealed that solar chimney as
a passive ventilation strategy has not been fully understood
despite continuous research eort devoted to the topic over
an extended period of time.
Bansal et al. (1993) established an idea of enhanced
stack ventilation using solar chimney based on a steady-
state mathematical model. A number of studies of this
topic are reported subsequently. Among these studies,
Bouchair (1994) considered the idea of nocturnal ventila-
tion using solar chimney for hot and arid climates in South-
ern Algeria through experimental and analytical studies.
Measurements of air temperature and velocity were carried
out for a full scale model under steady-state conditions.
The emphasis was placed on the eect of the cavity width
(air gap width) and the inlet aperture area on the mass ow
rate. Reverse ow was observed for a 0.5 m wide cavity.
http://dx.doi.org/10.1016/j.solener.2014.05.032
0038-092X/ 2014 Elsevier Ltd. All rights reserved.

Corresponding author. Tel.: +61 2 9351 5155; fax: +61 2 9351 3343.
E-mail address: r.khanal@sydney.edu.au (R. Khanal).
www.elsevier.com/locate/solener
Available online at www.sciencedirect.com
ScienceDirect
Solar Energy 107 (2014) 461474
Chen et al. (2003) and more recently, Jing and Li (2012)
also reported the occurrence of the reverse ow at the
chimney exit. Both of these experiments were performed
in an enclosed environment in which solar energy was sim-
ulated using an electrical heating system.
The occurrence of the reverse ow was also conrmed
numerically by several other researchers. Among these
investigations, Gan and Riat (1998) reported the occur-
rence of the reverse ow for a solar chimney with a 2.8 m
high absorber wall under an asymmetrical heating condi-
tion. They reported that, when the chimney width (i.e.
the air gap width) was increased beyond 0.2 m, the air ow
rate decreased, which was attributed to the eect of the
reverse ow. Further, for a wide chimney of 0.5 m air
gap width, the reverse ow was reported to penetrate
downwards at the centre of the chimney. A similar observa-
tion of the reverse ow was reported by Bouchair (1994)
for a 0.5 m wide asymmetrically heated solar chimney.
In a separate investigation, Gan (2006) reported the
occurrence of the reverse ow for a 6 m tall chimney with
an air gap width greater than 0.55 m under asymmetrical
heating conditions. The existence of the reverse ow was
also reported by Khanal and Lei (2011b) under an asym-
metrical heating condition for a small size solar chimney
with an absorber height less than 1 m and for various air
gap widths and inlet aperture heights.
All the above reported investigations were carried out
for asymmetrical heating conditions. For symmetrical heat-
ing conditions, Zamora and Kaiser (2009) also reported the
occurrence of the reverse ow at the exit of the channel at a
Rayleigh number of 10
8
for an aspect ratio (the ratio of the
air gap width to the absorber height) of 0.25 in their
numerical study. A considerable reduction of the mass ow
rate was attributed to the eect of the reverse ow.
Despite that the reverse ow has frequently been
observed and reported in both experimental and numerical
studies relevant to solar chimney ventilation, until recently
no detailed investigation of the reverse ow phenomenon
and its impact on the ventilation performance of solar
chimney has been reported.
In a recent numerical investigation, Khanal and Lei
(2012) attempted to quantify the reverse ow phenomenon
and its impact on the ventilation performance of a solar
chimney in terms of mass ow rate prediction. It was
reported that the reverse ow occurred as a result of
entrainment of air from downstream (chimney exit) into
the thermal boundary layer developing along the absorber
wall of the solar chimney, the penetration depth of which
increased monotonically with the Rayleigh number. The
occurrence of the reverse ow is not desirable for ventila-
tion applications. Therefore, in order to enhance the venti-
lation performance of a solar chimney for high Rayleigh
number applications, a new solar chimney design, namely
an inclined passive wall solar chimney (IPWSC), was pro-
posed by Khanal and Lei (2012). The eectiveness of this
design has been demonstrated numerically, which was sup-
ported by scale analysis and conrmed qualitatively by a
smoke ow visualization experiment.
After a further survey of the literature, it is revealed that
except for the qualitative experimental investigation men-
tioned above, no detailed experimental investigation of
the IPWSC has been reported in the public domain. In
addition, most of the reported studies involved large solar
chimneys. The study of small solar chimney for ventilation
Nomenclature
A aspect ratio
ACH Air Change per Hour (h
1
)
Bo Boussinesq number (=RaPr)
D
p
penetration depth of the reverse ow (m)
Gr
y
local Grashof number
Ra
Pr

g gravitational acceleration (m/s


2
)
H
a
absorber height (m)
k thermal conductivity (W/m K)
_
M
opt
optimum mass ow rate
Nu
y
local Nusselt number
q
00
Wy
y
DTk

Pr Prandtl number
t
a

q
00
input heat ux (W/m
2
)
R
a
global Rayleigh number
gbq
00
H
4
a
amk

Ra
y
local Rayleigh number
gbq
00
W
y
4
amk

T local uid temperature (K)
T
a
ambient air temperature (K)
T
bs
steady state temperature scale for the thermal
boundary layer (K)
T
W
wall temperature (K)
V vertical component of the velocity (m/s)
V
s
steady state velocity scale of the thermal bound-
ary layer (m/s)
W
g
air gap width (m)
x, y and z horizontal, vertical and spanwise coordinates
(m)
Greek symbols
a thermal diusivity (m
2
/s)
b coecient of thermal expansion (1/K)
DT temperature dierence from the ambient tem-
perature (K)
d
T
steady state thickness of the thermal boundary
layer (m)
h inclination angle of the passive wall ()
t kinematic viscosity (m
2
/s)
462 R. Khanal, C. Lei / Solar Energy 107 (2014) 461474
has received relatively less research attention despite its
potential for room ventilation (see for example Bansal
et al., 2005; Mathur et al., 2006; Chantawong et al.,
2006). These facts have motivated the present work.
In this work, experiments are carried out using an
IPWSC model with a uniform heat ux on the vertical
active wall under steady state conditions. The temperature
and air ow rates for dierent Rayleigh numbers and incli-
nation angles of the passive wall for a xed base air gap
width are measured and ow visualization experiments
with both smoke and shadowgraph techniques are carried
out. This is the rst attempt to visualize the convective ow
inside the solar chimney using the shadowgraph ow visu-
alization technique.
2. Experimental setup and procedure
Fig. 1 shows the schematic view of an inclined passive
wall solar chimney experimental system. The system is con-
structed to resemble the essential features of a solar chim-
ney directly attached to a ventilated space (room). It
consists of a transparent glazing wall at one side, opposite
to which is a heating plate for simulating the absorber wall
of the solar chimney; in between these two plates is the air
channel. The air channel is connected to the ventilated
space through the opening at the bottom of the heating
plate assembly. Fresh and cool air entering into the room
through the inlet window is drawn into the air channel hor-
izontally and exits the chimney vertically from the top.
The heating plate assembly of the solar chimney is com-
posed of two identical 3 mm thick aluminium plates,
700 mm high and 500 mm wide, between which a custom
made 1.4 mm thick breglass-reinforced silicon rubber hea-
ter (Watlow

) is sandwiched. Thus, the total thickness of


the heating plate assembly is 7.4 mm. The entire plate
assembly is then inserted into two vertical slots, which
are milled on the inside surface of the two vertical side
walls of the room, forming one surface of the air channel.
The other part of the air channel, i.e. the glazing wall is
made from a 6 mm thick, 800 mm high Perspex sheet. All
the other vertical walls of the room (enclosure) and the
channel are also made from 6 mm thick Perspex sheets
except for the top and bottom horizontal surfaces which
are made from 10 mm thick Perspex sheets. The glazing
wall is hinged at the bottom and attached to two threaded
rods. These rods are attached to a threaded cross bar,
which in turn is attached to the two vertical side walls of
the room (refer to Fig. 1). This arrangement gives extra
structural stability to the experimental rig.
The inclination of the glazing wall is altered by adjusting
the positions of the locking screws to tilt the glazing wall
inwards. The design of the apparatus allows both the base
air gap width and the inclination of the glazing wall (pas-
sive wall) to be adjusted so that the size of the exit aperture
changes accordingly. It is to note that this investigation
focuses on understanding the system behaviour of the
IPWSC under various Rayleigh numbers and inclination
angles of the passive wall. Other controlling parameters
such as the base air gap width are not considered. In addi-
tion, for a conventional solar chimney with 0.7 m absorber
height and 0.1 m inlet aperture size, the previous investiga-
tion of Khanal and Lei (2011) shows that the optimum
result could be achieved at the air gap width of 0.1 m.
We consider this as the reference point to size the base
air gap width of the present experimental model.
Electric power is supplied to the silicon rubber heater
through a variable auto transformer, which delivers a con-
stant power input to the rubber heater in order to maintain
a constant heat ux. This arrangement of heating to simu-
late constant solar heat ux is appropriate for the purpose
of the present experiment and is easier to control than
other arrangements such as articial lighting. Two 12 mm
1. Inlet window
2. Inside space (room)
3. Insulation
4. Heating plate assembly (absorber wall)
5. Glazing (passive wall)
6. Hinge
7. Threaded rods
8. Variable auto transformer
5
1
3
2
7
4
6
Wg
x
z
y
8

H
a
Exit air gap width
at y/H
a
= 1
Fig. 1. Schematic view of an inclined passive wall solar chimney experimental model.
R. Khanal, C. Lei / Solar Energy 107 (2014) 461474 463
thick insulation boards of high density material of the same
size as the heating plate are xed at the back side of the
heating plate assembly as shown in Fig. 1 to reduce the
heat loss to the room side. Insulation is also provided at
the bottom and the top of the heating plate assembly to
further reduce the heat loss.
The experimental rig is instrumented to measure the
temperature proles across the air gap width and along
the chimney height. A number of thermocouples (see
details below) are used for this purpose to measure the
absorber wall temperature and the air temperature in the
channel during the experiments. Fig. 2 presents a schematic
showing a typical arrangement of thermocouples in the
chimney channel for the zero degree inclination case.
Since most of the temperature variations (i.e. large tem-
perature gradients) are expected to occur near the absor-
ber, more thermocouples are placed in the vicinity of the
absorber wall. A total of 59 ne bead (with a diameter less
than 1 mm) K-type thermocouples, depending upon the
inclination of the glazing wall, are used for the temperature
measurement across the air gap width. A total of 15 K-type
thermocouples are used for the temperature measurement
along the chimney height at four positions across the air
gap width (refer to Fig. 2b). Inside the air channel, thermo-
couples are mounted on a small vertical rod (diameter
1 mm) which can be positioned at any desired location.
An additional thermocouple is used to monitor the ambi-
ent room temperature.
The thermocouples used for temperature measurements
across the air gap width and along the chimney height are
connected to two DaqPRO 5300 data loggers and are
calibrated in a constant temperature water bath (Julabo,
Model FP45) over the temperature range of 3080 C.
The uncertainty in the temperature measurement is esti-
mated to be 0.7 C following the guidelines on reporting
uncertainties in experimental measurements (Moat, 1988).
A typical experimental run is started at the ambient tem-
perature (so that the inlet temperature is always less than
the outlet air temperature) and the thermocouple readings
are veried before the system is heated up by recording the
temperature of the air for about 5 min. It is found that in
general the variations between the readings of the thermo-
couples do not exceed 0.2 C. Following the verication of
the thermocouple readings, the electric supply to the rubber
heater is switched on and the input power is adjusted to a
required level, which is monitored to ensure that it remains
constant throughout each test run. Depending upon the
input power supplied to the rubber heater, it takes about
34 h for the system to reach thermal equilibrium
(steady-state), during which the ambient air temperature
increases by up to 2 C.
For each set of the experiments, the heat ux on the
heating plate is xed at a constant value by adjusting the
power supply to the rubber heater, as described above.
The temperature measurements are then carried out over
a range of the inclination angles, starting from the zero
inclination (i.e. with the glazing wall at the vertical posi-
tion). Once the system has reached thermal equilibrium,
the temperature prole across the air gap width is mea-
sured rst, followed by the temperature prole along the
chimney height. After the temperature measurements for
a particular angle of inclination are nished, the inclination
Figure not to scale; Thermocouple positions indicated by x
(a)
Air Gap Width
W
i
d
t
h

o
f

t
h
e

c
h
a
n
n
e
l
Plan
Glazing
Heating plate
assembly
Air Gap Width
A
b
s
o
r
b
e
r


H
e
i
g
h
t
(b)
Elevation
Glazing
Heating plate
assembly
x
x
z
y
Fig. 2. Schematics showing the arrangement of the thermocouples (a) across the air gap width and (b) along the chimney height.
464 R. Khanal, C. Lei / Solar Energy 107 (2014) 461474
of the glazing wall is set to another value, and the same
temperature prole measurements are repeated. Apart
from the temperature measurements, the air velocity in
the channel is also measured at the steady-state. The air
velocity is measured using a robust hot-ball thermal
anemometer probe (Testo, model 06280035), which is con-
nected to a Testo 445 multifunctional measuring instru-
ment. The typical uncertainty of the velocity
measurement is 0.03 m/s.
The velocity probe is placed inside the chimney channel
at 560 mm above the leading edge. The corresponding
smoke ow visualization (refer to Section 3.5) showed that
at this position the disturbance from the reverse ow is
minimum for the highest Rayleigh number considered in
this study (Ra = 4.19 10
11
). In order to obtain the total
airow rate through the chimney, the probe is positioned
at various locations on the cross-sectional plane of the air
channel at the xed height (i.e. 560 mm above the leading
edge). The velocity readings are taken at spatial intervals
between 5 mm to 10 mm across the air gap width depend-
ing upon the inclination angle, and each reading is taken
for 60 s at a sampling rate of 1 Hz.
3. Experimental results and discussion
The present experiment is focused on studying small win-
dow-sized solar chimney design, and the model considered
here has an absorber height less than 1 m (0.7 m) with an
aspect ratio of 7 and a xed base air gap width of 0.1 m.
For most of the locations in Australia, an average solar irra-
diation of 100500 W/m
2
is common. Accordingly, in order
to study the performance of the IPWSC under the heat
input conditions corresponding to the above range of solar
irradiation, three Rayleigh numbers of 8.05 10
10
,
2.76 10
11
and 4.19 10
11
are considered with various
inclination angles ranging from 0 to 6.
3.1. Temperature distribution across the air gap width
Fig. 3 shows the non-dimensional temperature distribu-
tion across the air gap width measured at the location of
y/H
a
= 0.95 at various inclination angles and dierent
Rayleigh numbers. In this gure, the temperature is non-
dimensionalized by (T T
a
)/(T
W
T
a
) and the distance
from the absorber wall is normalised by the air gap width
(x/W
g
). The normalisation is for easy comparison with
the analytical solution, which will be presented later in this
section (refer to Fig. 4).
It is seen in Fig. 3 that the dimensionless temperature
distributions for dierent Rayleigh numbers and inclina-
tion angles remain fairly similar inside the chimney. As
expected, the temperature gradient is the highest near the
absorber wall and gradually diminishes away from the
absorber. It is also seen in Fig. 3 that the temperature dif-
ference decreases rapidly with the increase of the inclina-
tion angle and the temperature prole becomes very steep
at the largest inclination angle (i.e. for the 6 inclination).
However, the temperature distribution near the heated wall
is almost unaected by the inclination angle. This result
shows that the temperature structure within the thermal
boundary layer (in the region closest to the absorber wall)
is not aected by the inclination of the passive wall.
For natural convection along a vertical plate with a
prescribed heat ux, the temperature distribution in the
thermal boundary layer can be approximated by the fol-
lowing polynomial expression (Sparrow, 1955):
T T
a

q
00
d
T
2k
1
x
d
T

2
; 1
where
q
00
d
T
k
DT T
W
T
a
. Therefore, the temperature
distribution in a non-dimensional form can be expressed as
x/W
g
Inclination angle
(a)
x/W
g
Inclination angle
(b)
x/W
g
Inclination angle
(c)
(
T
-
T
a
)
/
(
T
W
-
T
a
)
(
T
-
T
a
)
/
(
T
W
-
T
a
)
(
T
-
T
a
)
/
(
T
W
-
T
a
)
Fig. 3. Dimensionless temperature distribution across the chimney air gap
width measured at y/H
a
= 0.95 for various inclination angles of the
passive wall and dierent Rayleigh numbers.
R. Khanal, C. Lei / Solar Energy 107 (2014) 461474 465
T T
a
T
W
T
a
0:5 1
x
d
T

2
: 2
The thermal boundary layer thickness can be approxi-
mated by the following expression (Rathore and Kapuno,
2011):
d
T
y
3:930:952 Pr
1=4
Gr
1=4
y
Pr
1=2
: 3
Fig. 4 shows the comparison between the dimensionless
temperature obtained for a vertical plate from Eq. (2) and
those from the present solar chimney experiment for two
dierent Rayleigh numbers, Ra = 4.19 10
11
and Ra =
2.76 10
11
. Here the temperature is measured in the exit
region of the solar chimney (i.e. at y/H
a
= 0.95) with 0
inclination of the passive wall. In this gure the abscissa
is the horizontal distance x perpendicular to the absorber
wall normalised by the base air gap width W
g
and the ordi-
nate is the normalised temperature dierence (T T
a
)/
(T
W
T
a
).
It is seen in Fig. 4a and b that close to the heated wall
the temperature proles for a vertical plate (the analytical
solution) and the solar chimney are very close to each
other. However, away from the heated surface, the experi-
mentally obtained air temperature in the solar chimney is
higher than the temperature given by the analytical solu-
tion, which rapidly reduces to the free stream value outside
the thermal boundary layer. The theoretical curve pre-
dicted by Eq. (3) depends on the thermal boundary layer
thickness and assumes that outside the thermal boundary
layer the air attains the undisturbed free stream tempera-
ture, whereas in the experiment the ambient temperature
increased by up to 2 C over the duration of the experi-
ment. The analytical solution does not account for the
ambient temperature variation. Similar observation of the
ambient temperature variation during experiment has been
reported by Ryan and Burek (2010).
3.2. Temperature distribution along the chimney height
Fig. 5a shows a typical temperature distribution along
the chimney height measured at dierent horizontal
positions from the heated surface with the 0 inclination
of the passive wall and the highest Rayleigh number con-
sidered in this study (Ra = 4.19 10
11
). It is to note that
at x = 0, i.e. at the absorber wall, the temperature of the
air is assumed to be the same as that of the wall and hence
the normalisation of the air temperature in the form of
(T T
a
)/(T
W
T
a
) as adopted above would always give
the same value of 1 at dierent positions on the absorber
wall. Therefore, a dierent normalisation scheme for the
temperature data is adopted here from that used in Sec-
tion 3.1. In Fig. 5, the ordinate is the temperature dierence
between the heated air and the ambient, (T T
a
), norma-
lised by the steady-state thermal boundary layer tempera-
ture scale, T
bs
q
00
H
a
=kBo
1=5
(Khanal and Lei, 2012),
and the abscissa is the distance from the leading edge, y
normalised by the absorber height, H
a
.
Fig. 4. Comparison of the dimensionless temperature distribution across the chimney air gap width measured at y/H
a
= 0.95 for 0 inclination angle with
the analytical result for (a) Ra = 4.19 10
11
and (b) Ra = 2.76 10
11
.
(a)
(b)
Fig. 5. (a) Dimensionless temperature distribution along the chimney
height measured at three dierent locations from the heated surface and
for Ra = 4.19 10
11
at 0 inclination. (b) Eect of the inclination angle on
the dimensionless temperature distribution along the chimney height
measured at three locations and for Ra = 4.19 10
11
.
466 R. Khanal, C. Lei / Solar Energy 107 (2014) 461474
It can be seen in Fig. 5a that the air temperature inside
the chimney increases along the height and the tempera-
ture prole becomes increasingly linear further away from
the absorber wall. It is also clear in this gure that the
temperature variation along the absorber wall shows an
unusual behaviour near the exit region. The temperature
on the absorber wall decreases as the exit is approached.
However, the temperature is expected to increase mono-
tonically in the streamwise direction. A similar tempera-
ture behaviour on the absorber wall has also been
reported by others (see for example Sandberg and
Moshfeg, 1996; Chen et al., 2003; Burek and Habeb,
2007; Jing and Li, 2012).
In a recent investigation, Jing and Li (2012) suggests
that this unusual behaviour near the exit could be due to
the relatively high heat transfer rate at the exit of the chim-
ney as a result of the transition to turbulence and further
enhancement of the heat transfer due to the increasing air
ow near the exit resulting from the occurrence of the
reverse ow.
As reported previously, the reverse ow near the exit of
the solar chimney can be minimized or even suppressed by
inclining the passive wall of the solar chimney (Khanal and
Lei, 2012). In order to understand the eect of the inclina-
tion of the passive wall on the temperature distribution
along the chimney height, especially the unusual behaviour
of the temperature prole on the absorber wall near the
exit, further experiments are carried out with dierent incli-
nation angles. The results are compared with the case of
zero degree inclination in Fig. 5b. It is clear from Fig. 5b
that the temperature distribution at all the three locations
is hardly aected by the inclination of the passive wall.
Since the inclination of the passive wall has a signicant
impact on the strength and extent of the reverse ow
(Khanal and Lei, 2012), this result suggests that the unu-
sual temperature behaviour observed in Fig. 5a is not due
to the reverse ow.
In an experiment with an asymmetrically heated vertical
channel, Sparrow et al. (1984) reported that the heat trans-
fer coecient is insensitive to the presence of the reverse
ow. This further suggests that the unusual temperature
behaviour observed near the exit region on the absorber
wall is not due to the occurrence of the reverse ow.
Bejan (1993) reported that the transition to turbulence
for a vertical plate heated by a uniform heat ux occurs
at Rayleigh numbers of the order of Ra
y
10
13
.
Furthermore, Chen et al. (2003) reported that the transi-
tion to turbulence occurs over the Rayleigh number range
of 2 10
13
< Ra < 1 10
14
for a vertical plate heated by a
uniform heat ux. In the present solar chimney experiment,
the highest Rayleigh number is Ra = 4.19 10
11
. There-
fore, the transition to turbulence is not expected to occur
in the present experiment. This will be conrmed by shad-
owgraph ow visualization described below (refer to
Section 3.4). It is hypothesized that if the transition were
to occur, then the shadowgraph ow visualization should
be able to capture certain turbulent ow behaviour.
3.3. Average air ow velocity and ACH prediction
Fig. 6 shows the dependence of the average air ow
velocity across the chimney air gap width on the inclination
angle of the passive wall for dierent Rayleigh numbers
with a xed base air gap width of 100 mm. In this gure,
the ordinate is the air ow velocity normalised by the
steady-state velocity scale, V
s
Bo
2/5
(a/H
a
) (Khanal and
Lei, 2012) and the abscissa is the inclination angle of the
passive wall.
It is to note that each point in Fig. 6 represents an aver-
aged velocity measured at ve dierent spatial locations
across the air gap width at y/H
a
= 0.8. It is seen from
Fig. 6 that, for all the three Rayleigh numbers tested here,
initially the averaged velocity increases with the increase of
the inclination angle but decreases beyond a certain incli-
nation angle. This increase of the air ow velocity inside
the chimney channel with the increase of the inclination
angle of the passive wall is attributed to the reduction of
the reverse ow occurring at the exit of the chimney. With
the reduction of the reverse ow, the thermal boundary
layer entrains more air from upstream; as a consequence,
the through ow becomes stronger. The increase of the
air ow velocity with the reduction of the reverse ow asso-
ciated with the inclination of the passive wall corresponds
well with the relatively stronger exit ow observed in
Figs. 11 and 14 (refer to Sections 3.4 and 3.5).
Fig. 6. Normalised air ow velocity plotted against the inclination angle
for dierent Rayleigh numbers and a xed base air gap width of 100 mm.
Fig. 7. Normalised mass ow rate plotted against the inclination angle.
R. Khanal, C. Lei / Solar Energy 107 (2014) 461474 467
With the increase of the inclination angle of the passive
wall, the exit air gap width reduces in size (refer to Figs. 11
and 14). Further increase of the inclination angle would
disturb the thermal boundary layer. Under this situation,
the air ow may be choked, resulting in reduction of the
mass ow rate (Khanal and Lei, 2012). This suggests that
for a given Rayleigh number, a maximum mass ow rate
exists at a certain inclination angle, beyond which the mass
ow rate reduces and so does the ventilation performance
(refer to Fig. 9). The increase of the mass ow rate with
the increase of the inclination angle and the reduction in
the mass ow rate beyond an optimum value of the inclina-
tion angle are also shown in Fig. 7. In this gure, the nor-
malised mass ow rate obtained experimentally for the
Rayleigh number of Ra = 4.19 10
11
is plotted against
the inclination angle of the passive wall. In the same gure,
the numerical data obtained for Ra = 10
11
, the highest
Rayleigh number considered in Khanal and Lei (2012),
are also plotted. Here the mass ow rate is normalised by
the optimum (maximum) mass ow rate in each case. It
is clear from the gure that both the experimental and
numerical results exhibit a similar trend; however, the opti-
mum mass ow rates are dierent between the experiment
and simulation. The variation between the experiment and
simulation is mainly due to the dierent Rayleigh numbers
adopted in these investigations, which results in dierent
optimum inclination angles. Nevertheless, both the experi-
mental and numerical results conrm that, for a given Ray-
leigh number, there exists an optimum inclination angle at
which the mass ow rate is the maximum (Khanal and Lei,
2012).
For Rayleigh numbers of 8.05 10
10
, 2.76 10
11
and
4.19 10
11
the maximummass owrate is achieved at incli-
nation angles of 3, 4 and 6 respectively (refer to Fig. 9).
This experimental result clearly demonstrates that the
IPWSC is an eective design for improving the ventilation
performance of a solar chimney in comparison to the con-
ventional design of a vertical passive wall conguration.
The maximum (optimum) mass ow rate for a solar
chimney can be predicted from the following scaling rela-
tion (Khanal and Lei, 2012)
_
M
opt
Bo
1=5
4
Fig. 8 plots the experimental data against the scaling
prediction of (4). This gure clearly shows a linear depen-
dence between the experimentally measured mass ow rate
and that of the scaling prediction, which conrms the
dependence of the optimum mass ow rate on the
Boussinesq number. Since air is the working uid in solar
chimney whose Prandtl number is approximately constant
(Pr 0.71), the mass ow rate can be expressed as a func-
tion of the Rayleigh number only.
To put these results into a more practical perspective, if
one considers an average solar irradiation of 500 W/m
2
,
which corresponds to the maximum Rayleigh number con-
sidered in this study and is common in most locations in
Australia, a 0.7 m tall IPWSC with 6 inclination and a
base air gap width of 100 mm yields a volumetric ow rate
of 0.0186 m
3
/s. If this ventilation rate is applied to a 27 m
3
room, the ACH is found to be around 2.5. This is more
than the adequate ventilation rate according to the stan-
dards ASHREA 66.2-2003 (Sherman, 2003) and ASHREA
62-2001 (Standard, 2001).
Further to the above discussion, Fig. 9 plots the ACH
that can be achieved if natural ventilation is considered
for a 27 m
3
room using the IPWSC at dierent Rayleigh
numbers. In this gure the dotted horizontal line indicates
the adequate rate of Air Change per Hour based on the
ASHREA 66.2-2003 standard. Fig. 9 shows that with the
vertical passive wall conguration of the conventional solar
chimney design, the generated ventilation rate in terms of
the ACH is lower than the recommended value at the low-
est Rayleigh number (Ra = 8.05 10
10
). On the other
hand, this recommended value of ACH is achieved with
the inclined passive wall conguration even for the lowest
heat input case. Therefore, with the IPWSC conguration
adequate ventilation can be achieved even for small heat
input corresponding to low solar irradiation. It is to note
that in Fig. 9, the data for the highest Rayleigh number
case (Ra = 4.19 10
11
) is extended to 6.5 inclination to
show the decreasing trend in the ACH after reaching the
optimum value.
It is seen in Fig. 9 that the maximum ventilation rates
achieved with the IPWSC design for high Rayleigh
numbers are signicantly greater than the recommended
Bo
1/5
opt
M
Fig. 8. Normalised mass ow rate plotted against scaling prediction.
Fig. 9. Ventilation rate plotted against the inclination angle of the passive
wall for dierent Rayleigh numbers and for a xed base air gap width of
100 mm.
468 R. Khanal, C. Lei / Solar Energy 107 (2014) 461474
standard. For Ra = 8.05 10
10
, the maximum ventilation
rate is achieved at 3 inclination and this value is still 9%
higher than the recommended standard. Similarly the max-
imum ventilation rate is achieved at 4 and 6 for Ra = 2.76
10
11
and Ra = 4.19 10
11
and is 63% and 90% higher
than the recommended value respectively. This is a
signicant improvement. Therefore, the IPWSC design is
very eective in enhancing natural ventilation for high
Rayleigh number (corresponding to high solar irradiation)
applications.
3.4. Shadowgraph ow visualization
Visualization of natural convection ows using the
shadowgraph technique with water as a working uid has
been widely reported (see for example Scho pf and
Patterson, 1995; Schopf and Patterson, 1996; Lei and
Patterson, 2002; Xu et al., 2006). However, the use of the
shadowgraph technique for visualizing natural convection
ows in air is rarely reported, and to the best of the
authors knowledge, shadowgraph ow visualization of
natural convection in solar chimney has not been reported
in the open literature. An attempt has been made in this
study to visualize the natural convection ow in a solar
chimney using the shadowgraph technique under steady-
state conditions in order to obtain further insight into the
ow behaviour, and in particular, to conrm if the transi-
tion to turbulence occurs inside the solar chimney.
The shadowgraph technique is based on the principle of
optical refraction. The distribution of the light intensity in
the shadowgraph image represented the spatial tempera-
ture eld inside the uid (for more detailed explanation
refer to Merzkirch, 1987; Lei and Patterson, 2002).
The optical setup used in the present research for shad-
owgraph visualization is schematically depicted in Fig. 10.
A point light source is generated by placing a pinhole in
front of a halogen lamp, which is placed at a distance of
2.4 m away from the rst spherical mirror (i.e. on the focus
point of the mirror). The diameter of the mirror is 30 cm.
The resulting parallel light beam passes through the
experimental model in such a way that the working region
of interest is illuminated. A second mirror, which is identi-
cal to the rst one, converges the transmitted light onto a
CCD-camera. The images are then recorded in a computer
at a rate of 5 frames per second, which are subsequently
processed and analysed. It should be noted that, since the
diameter of the spherical mirrors (30 cm) is smaller than
the total chimney height (0.8 m), the entire chimney height
cannot be displayed. Accordingly, only the region of the
most interest is displayed, i.e. the exit region and the region
above the outlet. It is to note that the shadowgraph tech-
nique is a simple, inexpensive and powerful method of ow
visualization, yet it is very sensitive. Therefore, utmost care
is needed during the set-up and experiment.
Fig. 11 shows the shadowgraph images captured for a
representative case of Ra = 4.19 10
11
. In this gure, for
clarity, the position of the heating plate assembly is indi-
cated by the red dashed line. It is seen in Fig. 11 that air
is being ejected like a buoyant jet from the chimney and
this ow is concentrated in the region near the absorber
wall, where the temperature gradient is the highest. Also
a bright strip is seen to form next the heated wall which
is approximately parallel to the absorber wall over the
observation region. Similar observation regarding the for-
mation of bright strips adjacent to heat transfer surfaces
is reported by Xu et al. (2008). The bright strip approxi-
mately represents the edge of the thermal boundary layer.
The boundary layer thickness is thus considered as the
width of the dark region between the heating plate (as indi-
cated by the dashed line) and the bright strip. The dark
region is formed as a result of light being deected away
from this region and indicates the existence of strong non-
linearity of the temperature gradient in this region since
light deection is roughly sensitive to the second spatial
derivatives of the temperature (Merzkirch, 1987). The
bright strip is caused by light being deected to this region
and corresponds to the minima of the second derivative of
the temperature (Xu et al., 2008). In the region between
the bright strip and the glazing wall, the shadowgraph
appears approximately the same for all inclination angles,
1
2
3
4
1. Point light source
2. Spherical mirror
3. CCD-Camera
4. Experimental model
2
4
Fig. 10. Schematic of the optical set-up for the shadowgraph visualization.
R. Khanal, C. Lei / Solar Energy 107 (2014) 461474 469
indicating that the temperature gradient in this region is
approximately constant. The temperature distribution pre-
sented in Section 3.1 is consistent with this observation.
The shadowgraph images in Fig. 11 do not show any
erratic ow behaviour inside the chimney, especially near
the exit. Furthermore, the images do not show any sign
of turbulence occurring in the thermal boundary layer.
Otherwise, an approximately parallel bright strip adjacent
to the absorber wall would not be observed. These shadow-
graph images indicate that the transition to turbulence has
not occurred in the present case. Furthermore, an interest-
ing ow feature (see below) is observed at the exit of the
chimney as the inclination angle increases.
It is seen in Fig. 11 that the buoyant ow at the exit of
the chimney near the heated wall (above) becomes thicker
as the inclination angle increases. This behaviour is consis-
tent with the increase of the air ow velocity with the incli-
nation of the passive wall, which reduces the reverse ow
occurring at the chimney exit. This observation corre-
sponds well with the results presented in Fig. 6. It is note-
worthy that the reverse ow cannot be observed from the
shadowgraph visualization as the reverse ow is from the
ambient air and there is no appreciable temperature gradi-
ent in the ow to inuence the light deection.
Further to these observations, an attempt has been
made to calculate the heat transfer coecient for the case
of the highest Rayleigh number studied here. For a laminar
natural convection ow, the local Nusselt number can be
scaled as Nu
y
/ Ra
1=5
y
(Bejan, 1993).
It is to note that the local Nusselt number can be
expressed as Nu
y
y=d
T
, which follows from the relations
for the local Nusselt number, Nu
y
q
00
Wy
y=DTk and the local
heat ux, q
00
Wy
kDT=d
T
. Therefore, if the boundary layer
thickness d
T
is known for a given position, the local Nusselt
number can be evaluated accordingly. The boundary layer
thickness is estimated using the shadowgraph images and
the thickness of the thermal boundary layer is considered
as the width of the dark strip as discussed above. The sim-
ilar method for estimating the boundary layer thickness has
been previously employed by Lei and Patterson (2002) and
Xu et al. (2005).
Fig. 12 plots the local Nusselt number as a function of
the local Rayleigh number for the case of Ra = 4.19
10
11
. From Fig. 12, it is seen that the experimental data
obeys Nu
y
/ Ra
1=5
y
until near the exit region where it devi-
ates slightly from the linear correlation. The reason for this
slight deviation is not clear.
3.5. Smoke ow visualization
In conjunction with the shadowgraph ow visualization,
smoke ow visualization has also been carried out to reveal
further insight into the ow behaviour, especially the
reverse ow occurring near the exit of the chimney, which
cannot be observed from the shadowgraph ow visualiza-
tion. It is to note that a similar smoke ow visualization
experiment has been reported previously by Khanal and
Lei (2012), but no quantitative information was obtained
from the previous experiment. In the present work, the
smoke ow visualization is carried out over a range of incli-
nation angles at a higher Rayleigh number (4.19 10
11
)
compared with 10
10
in the previous experiment), and quan-
titative information regarding the penetration depth of the
reverse ow will be presented.
With the experimental model described in Section 2, the
set-up of the smoke ow visualization experiment is shown
in Fig. 13 for a xed base air gap width (100 mm) with
dierent inclination angles varying from 0 to 6. In the
present case, steady-state is achieved after around 3 h from
the start of the experiment.
After the attainment of the steady-state, ow visualiza-
tion is carried out. Smoke is pre-generated using an Antari
Fog machine and collected in a storage chamber made
from cardboard boxes (1600 mm 635 mm 635 mm).
This arrangement is to minimize the disturbance to the
convective ow caused by the smoke generator, which is
turned o during the smoke visualization experiment.
The stored smoke in the storage chamber is drawn natu-
rally into ventilated space which is connected to the air
channel through a cylindrical passage. The air movement
in the air channel is then photographed using a high speed
camera and images are stored at a rate of 10 frames per sec-
ond. The experiment is rst carried out with 0
o
inclination
angle and the same procedure is then repeated for the other
inclination angles (16).
The results of the smoke ow visualization for various
inclination angles of the passive wall are shown in
Fig. 14. This gure clearly shows the convective ow in
the channel between the heating plate assembly on the left
and the glazing wall on the right, the latter of which is
attached to the threaded rods.
Since the main objective of the smoke ow visualization
is to observe the reverse ow occurring near the exit of the
chimney, results are presented only for the region of inter-
est (i.e. near the exit region) but not for the entire chimney
channel. Fig. 14 shows more than half of the channel
height including the region immediately above the chimney
exit. It is clear from the gure that the ow at the chimney
exit becomes stronger with the increase of the inclination
angle of the passive wall, as indicated by the variation of
the jet ow from the exit. Also it can be observed that
the reverse ow is occurring at the exit of the channel
(for 02). For clarity, the region of the reverse ow is
approximately indicated by the dotted box in Fig. 14. This
region of reverse ow can be identied by observing the
smoke density (indicated by the colour or grey level of
the images) which appears to be relatively lighter in the
region with reverse ow due to the entrainment of the fresh
air from outside.
It is also seen in Fig. 14 that the region of the reverse
ow reduces with the inclination of the glazing wall. As a
consequence, the ow rate increases. This result supports
the above discussion in Section 3.2 and conrms that the
unusual temperature structure near the exit (refer to
470 R. Khanal, C. Lei / Solar Energy 107 (2014) 461474
Fig. 5a) is not caused by the reverse ow. In order to obtain
quantitative information about the reverse ow, images are
further processed and the penetration depth of the reverse
ow is estimated. Since the region of the reverse ow
appears relatively lighter in colour as a result of the
entrainment of fresh air, the penetration depth of the
reverse ow is estimated based on the height of the lighter
coloured region measured downwards from the chimney
exit, as indicated by the dotted lines in Fig. 14. The
limitation of the penetration depth measurement described
here is worth noting.
Fig. 15 plots the calculated penetration depth against
the inclination angle of the passive wall for Ra = 4.19
10
11
. Here the penetration depth is normalised by the max-
imum penetration depth. On the same gure, the numeri-
cally calculated penetration depth reported in Khanal
and Lei (2012) for Ra = 10
10
is also plotted. It is clear from
this gure that both the experimental and numerical data
exhibits the similar trend, showing an inverse correlation
of the penetration depth with the inclination angle of the
passive wall. It is also clear from this gure that for both
cases, the maximum penetration depth occurs at 0
inclination.
The penetration depth of the reverse ow can be pre-
dicted using the following scaling relation (Khanal and
Lei, 2012):
D
p
H
a

1
Pr
1
A

1
Bo
1=5

5
Fig. 16 plots the estimated penetration depth against the
scaling prediction (5). It is to note that in this gure, the
aspect ratio A is calculated as the ratio of the absorber
height to the exit aperture size. Therefore, an increase of
the inclination angle corresponds to a reduction of the exit
air gap width and thus an increase of the aspect ratio. It is
0
o
inclination 1
o
inclination 2
o
inclination 3
o
inclination
4
o
inclination 5
o
inclination
6
o
inclination
(d) (c)
(e) (g) (f)
(b) (a)
Fig. 11. Shadowgraph images showing the convective ow in an IPWSC. The images are for the steady state case with various inclination angles and
Ra = 4.19 10
11
. The dashed red line in (a) indicates the position of the absorber surface. (For interpretation of the references to colour in this gure
legend, the reader is referred to the web version of this article.)
Fig. 12. Local Nusselt number plotted against the local Rayleigh number.
The case is for 0
o
inclination with the global Rayleigh number
Ra = 4.19 10
11
.
R. Khanal, C. Lei / Solar Energy 107 (2014) 461474 471
clear from this gure that the experimental data agrees well
with the scaling prediction. Since the Boussinesq number is
xed for the case presented, the linear correlation between
the experimental and scaling data conrms the dependence
of the penetration depth on the aspect ratio which in turn
depends on the inclination angle.
Fig. 13. Schematic of the experimental set-up for smoke ow visualization.
Fig. 14. Results of smoke ow visualization experiment for Ra = 4.19 10
11
at dierent inclination angles of the passive wall.
472 R. Khanal, C. Lei / Solar Energy 107 (2014) 461474
4. Conclusions
In this study the buoyancy induced convective ow in an
inclined passive wall solar chimney (IPWSC) model is
experimentally studied with a uniform heat ux on the ver-
tical absorber wall for dierent Rayleigh numbers and
inclination angles of the passive wall while the base air
gap width is xed. For the IPWSC design, whilst the ow
velocity is found to be strongly aected by the inclination
of the passive wall, the temperature distribution is found
to be insensitive to the inclination of the passive wall. In
addition, the present temperature data conrmed an unu-
sual temperature behaviour at the exit of the chimney.
The evidences from the direct temperature measurements
and the shadowgraph and smoke ow visualizations at dif-
ferent inclination angles of the passive wall suggest that the
unusual temperature drop near the exit is not caused by the
reverse ow or transition to turbulence. Further investiga-
tions are required to resolve this issue.
It has been demonstrated experimentally that by inclin-
ing the passive wall of the solar chimney while maintaining
all the other conditions unchanged, the air ow rate
increases with the increase of the inclination angle and
reaches a maximum value at a certain inclination, beyond
which the air ow rate decreases. The increase of the air
ow rate with the increase of the inclination angle is attrib-
uted to the reduction of the reverse ow occurring near the
exit of the chimney.
The experimental results have showed that the IPWSC
design is eective for enhancing natural ventilation not
only for high Rayleigh number applications but also for
relatively low Rayleigh number applications. The present
experimental results have also showed that the IPWSC
can signicantly improve the ventilation rate in compari-
son to the conventional solar chimney design of the vertical
passive wall conguration. It is shown that for the highest
Rayleigh number considered in this study (Ra = 4.19
10
11
), the Air Change per Hour achieved for a room of
27 m
3
is substantially higher than the required Air Change
per Hour based on ventilation standards. The experimental
results are supported by shadowgraph and smoke ow
visualizations. The experiment data have also been com-
pared against the scaling prediction and numerical simula-
tion, which shows consistent behaviour.
Acknowledgements
Authors would like to thank Mr. Theo Gresley-Daines
and Mr. Langga Thepolaan for their assistance with the
experiment.
References
Bansal, N.K., Mathur, J., Mathur, S., Jain, M., 2005. Modeling of
window-sized solar chimneys for ventilation. Build. Environ. 40, 1302
1308.
Bansal, N.K., Mathur, R., Bhandari, M.S., 1993. Solar chimney for
enhanced stack ventilation. Build. Environ. 28, 373377.
Bejan, A., 1993. Heat Transfer. John Wiley & Sons Inc.
Bouchair, A., 1994. Solar chimney for promoting cooling ventilation in
Southern Algeria. Build. Services Eng. Res. Technol. 15, 8193.
Burek, S.A.M., Habeb, A., 2007. Air ow and thermal eciency
characteristics in solar chimneys and Trombe Walls. Energy Build.
39, 128135.
Chantawong, P., Hirunlabh, J., Zeghimati, B., Khedari, J., Teekasap, S.,
Win, M.M., 2006. Investigation on thermal performance of glazed
solar chimney walls. Sol. Energy 80, 288297.
Chen, Z.D., Bandopadhayay, P., Halldorsson, J., Byrjalsen, C., Heisel-
berg, P., Li, Y., 2003. An experimental investigation of a solar chimney
model with uniform wall heat ux. Build. Environ. 38, 893906.
Gan, G., 2006. Simulation of buoyancy-induced ow in open cavities for
natural ventilation. Energy Build. 38, 410420.
Gan, G., Riat, S.B., 1998. A numerical study of solar chimney for natural
ventilation of buildings with heat recovery. Appl. Therm. Eng. 18,
11711187.
Jing, H., Li, A., 2012. Experimental study of a vertical channel solar
chimney with unifrom heat ux for natural ventilation in buildings.
Adv. Mater. Res. 374377, 585589.
Khanal, R., Lei, C., 2011a. Solar chimney a passive strategy for natural
ventilation. Energy Build. 43, 18111819.
Khanal, R., Lei, C., 2011b. A numerical investigation of the ventilation
performance of a solar chimney. ANZIAM J. 52, C899C913.
Khanal, R., Lei, C., 2012. Flow reversal eects on buoyancy induced air
ow in a solar chimney. Sol. Energy 86, 27832794.
Lei, C., Patterson, J.C., 2002. Natural convection in a reservoir sidearm
subject to solar radiation: experimental observations. Exp. Fluids 32,
590599.
Fig. 15. Normalised penetration depth plotted against the inclination
angle for Ra = 4.19 10
11
(experiment) and Ra = 10
10
(simulation).
Fig. 16. Normalised penetration depth plotted against scaling prediction
for Ra = 4.19 10
11
.
R. Khanal, C. Lei / Solar Energy 107 (2014) 461474 473
Mathur, J., Bansal, N.K., Mathur, S., Jain, M., Anupama, 2006.
Experimental investigations on solar chimney for room ventilation.
Sol. Energy 80, 927935.
Merzkirch, W., 1987. Flow Visualization. Academic Press Inc., Ltd.
Moat, R.J., 1988. Describing the uncertainties in experimental results.
Exp. Thermal Fluid Sci. 1, 317.
Rathore, M.M., Kapuno, R.R., 2011. Engineering Heat Transfer. Jones &
Bartlett Learning LLC.
Ryan, D., Burek, S.A.M., 2010. Experimental study of the inuence of
collector height on the steady state performance of a passive solar air
heater. Sol. Energy 84, 16761684.
Sandberg, M., Moshfeg, B., 1996. Investigation of uid ow and heat
transfer in a vertical channel heated from one side by PV elements,
Part II Experimental study. Renew. Energy 8, 254258.
Scho pf, W., Patterson, J., 1996. Visualization of natural convection in a
side-heated cavity: transition to the nal steady state. Int. J. Heat Mass
Transf. 39, 34973509.
Scho pf, W., Patterson, J.C., 1995. Natural convection in a side-heated
cavity: visualization of the initial ow features. J. Fluid Mech. 295,
357380.
Sherman, M., 2003. ASHRAEs rst residential ventilation standard.
LBNL Report Number.
Sparrow, E., 1955. In: Lewis Flight Propulsion Laboratory, C. (Ed.),
Laminar Free Convection on a Vertical Plate With Prescribed Non
Uniform Wall Heat Flux or Prescribed Non Uniform Wall Temper-
ature. National Advisory Committee for Aeronautics, Ohio.
Sparrow, E.M., Chrysler, G.M., Azevedo, L.F., 1984. Observed ow
reversals and measured-predicted nusselt numbers for natural convection
in a one-sided heated vertical channel. J. Heat Transfer 106, 325332.
Standard 62-2001, 2001. Ventilation for Acceptable Indoor Air Quality.
American Society of Heating, Refrigerating, and Air-Conditioning
Engineers, Atlanta, <www.ASHRAE.org>.
Xu, F., Patterson, J.C., Lei, C., 2005. Shadowgraph observations of the
transition of the thermal boundary layer in a side-heated cavity. Exp.
Fluids 38, 770779.
Xu, F., Patterson, J.C., Lei, C., 2006. Experimental observations of the
thermal ow around a square obstruction on a vertical wall in a
dierentially heated cavity. Exp. Fluids 40, 364371.
Xu, F., Patterson, J.C., Lei, C., 2008. On the double-layer structure of the
boundary layer adjacent to a sidewall of a dierentially heated cavity.
Int. J. Heat Mass Transf. 51, 38033815.
Zamora, B., Kaiser, A.S., 2009. Optimum wall-to-wall spacing in solar
chimney shaped channels in natural convection by numerical investi-
gation. Appl. Therm. Eng. 29, 762769.
474 R. Khanal, C. Lei / Solar Energy 107 (2014) 461474

Você também pode gostar