You are on page 1of 9

Heat and Mass Transfer Analysis

Ramendra Pandey
Academy of Scientific and of a Pot-in-Pot Refrigerator
Innovative Research (AcSIR),
CSIR-National Chemical
Laboratory (CSIR-NCL),
Using Reynolds Flow Model
Dr. Homi Bhabha Road,
Pashan, Pune 411008, India Heat and mass transfer analysis of evaporative cooling process in a pot-in-pot cooling sys-
e-mail: ramendra.csir@gmail.com tem is done based on Reynolds flow hypotheses. The model proposed herein assumes that
the heat transfer due to natural convection is coupled with an imaginary ambient air mass
Bala Pesala flow rate (gAo) which is an essential assumption in order to arrive at the solution for the
Academy of Scientific and rate of water evaporation. Effect of several parameters on the pot-in-pot system perform-
Innovative Research (AcSIR), ance has been studied. The equations are iteratively solved and detailed results are pre-
CSIR-Structural Engineering Research sented to evaluate the cooling performance with respect to various parameters: ambient
Centre (CSIR-SERC) Campus, temperature, relative humidity (RH), pot height, pot radius, total heat load, thermal and
Chennai 600113, India; hydraulic conductivity, and radiation heat transfer. It was found that pot height, pot radius,
total heat load, and radiation heat transfer play a critical role in the performance of the
CSIR-Central Electronics Engineering
system. The model predicts that at an ambient temperature of 50  C and RH of 40%, the
Research Institute,
system achieves a maximum efficiency of 73.44% resulting in a temperature difference of
CSIR Madras Complex,
nearly 20  C. Similarly, for a temperature of 30  C and RH of 80%, the system efficiency
Chennai Unit,
was minimum at 14.79%, thereby verifying the usual concept that the pot-in-pot system is
Taramani, Chennai 600113, India
best suited for hot and dry ambient conditions. [DOI: 10.1115/1.4033010]
e-mail: balapesala@gmail.com
Keywords: evaporative cooling, heat transfer, mass flux, pot-in-pot, Reynolds flow

1 Introduction air–water–clay interfaces. Convection and radiation effects of the


ambient atmosphere provide the energy to the surface water mole-
A pot-in-pot refrigerator is a simple to fabricate and a cost
cules of the outer pot to evaporate faster [1], which causes cooling
effective means of food storage. It primarily uses two clay pots in
at the surface. In addition, the ambient causes heat load to the
which one fits inside the other and the annular space is filled with
inner pot through conduction. While convection and radiation
sand. Food is stored inside the inner pot. Figure 1 shows a sche-
heat transfer are dominant over the outer pot surface, the heat
matic illustration for building a pot-in-pot system. The bigger pot
transfer from inner pot to the outer pot is achieved mainly due to
should be big enough to accommodate the smaller pot with
conduction. The performance of the system is primarily affected
2–3 cm annular space for sand. Once the pots are set with sand
by ambient temperature, RH, and several material and geometrical
filled in the annular space, water is filled up to the saturation point.
properties of the system [2–4]. The minimum temperature that the
The top of the pot is covered by a wet cloth or any thermal insulat-
outer pot surface can reach through passive evaporation is the wet
ing material. Evaporative cooling is achieved in the system
because of the wet sand that is present in the annular space. The
water present in the sand seeps through the pores of the outer clay
walls and evaporates into the ambient atmosphere. Because of this
evaporation process, cooling is achieved in the inner pot.
In many rural areas of the world, people make use of pot-in-pot
refrigerator for preserving their fruits and vegetables. The system
is also ideally suited for vendors who earn their living by selling
fruits and vegetables under hot and dry climatic conditions. Sev-
eral thousands of villages in most parts of under developed and
developing countries, accounting to nearly 20% of the world’s
population, still do not have access to electricity. Depending upon
the storage infrastructure in the individual countries, the annual
loss of fruits and vegetables due to lack of proper storage accounts
to about 35–50%. Hence, there is a strong requirement for passive
systems that can store vegetables/fruits for longer durations. Com-
pared to the conventional refrigeration which runs on electricity
and is costly, frugal innovations such as pot-in-pot systems can be
a boon for places without access to electricity and for people who
cannot afford costly conventional refrigerators.
The physical principle behind the operation of pot-in-pot sys-
tem is evaporation of water from the wet porous clay surface into
the ambient dry air stream. The cooling effect is produced due to
the coupled heat and mass transfer processes that occurs at the

Contributed by the Heat Transfer Division of ASME for publication in the


JOURNAL OF THERMAL SCIENCE AND ENGINEERING APPLICATIONS. Manuscript received
September 6, 2015; final manuscript received February 9, 2016; published online
April 12, 2016. Assoc. Editor: Amir Jokar. Fig. 1 Schematic illustration for building a pot-in-pot system

Journal of Thermal Science and Engineering Applications SEPTEMBER 2016, Vol. 8 / 031006-1
Copyright VC 2016 by ASME

Downloaded From: http://thermalscienceapplication.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jtsebv/935128/ on 03/05/2017 Terms of Use: http://www.as


bulb temperature. However, if the moist air of the enclosed space
of the inner pot is cooled under constant pressure conditions, then
the moisture will begin to condense at the dew point temperature.
We, therefore, assume that this cooling device can achieve this
ideal state of cooling. Therefore, the effectiveness of such a refrig-
eration system is given by the following equation:

T1  Tcold
g¼ (1)
T1  Tdp

where T1 is the ambient temperature, Tcold is the cold zone tem-


perature, and Tdp is the dew point temperature which is the mini-
mum attainable temperature due to thermodynamic limitations. If
we consider the convection and radiation effects of the ambient
air as the work done on the system, we can also determine the
steady-state coefficient of performance (COP) for such a system,
which can be given as the ratio of respiratory heat load to the
work input, i.e., ambient heat. Thus, it can be expressed as

Q_ load q_  mfood
COP ¼ ¼ load (2)
_
Q in _
Q rad þ gAo h1
Fig. 2 Different states in Reynolds flow model
where Q_ load accounts for the respiratory heat load for storing mfood
(kg) amount of food and Q_ in stands for the input energy that is
taken up by the system from the ambient. the Reynolds flow model for a pot-in-pot system. The model is
Although this cooling device has been in existence for a very assumed symmetric about the radial axis with the base and top of
long time in the history, very few attempts have been made to- the pot-in-pot system insulated.
ward modeling of the device and quantification of the system’s In this model, the transferred substance is taken at uniform
behavior. This is mainly due to simultaneous multiple processes temperature and concentration throughout. However, since the
occurring in the system, which depend on several material and temperature in the annular space occupied by the transferred sub-
geometrical parameters. Moreover, comprehensive modeling of stance will vary, we construct an imaginary transferred substance
such processes is extremely complicated because of various multi- state designated as M–M with uniform properties given by
physics effects linked with this phenomenon. Date developed a
mathematical model that accounts for the coupling of heat and Ti þ To
TM ¼ and xv;M ¼ 1
mass transfer in the clay pot refrigerator using a simplified alge- 2
braic model based on Reynolds flow hypotheses [5]. However, the where Ti and To are temperatures of the outer surface of the inner
energy balance equation used therein fails to provide the solution pot and inner surface of the outer pot, respectively. xv;M is the
for the evaporation rate of water and, therefore, has limitations in mass fraction of water vapor in the transferred substance state.
correctly predicting the steady-state temperature and effectiveness Reynolds flow model is based on the following hypotheses:
of the system with respect to varying ambient conditions and
material parameters. The heat balance equation used in Ref. [5] (1) a mass flow rate of (gAo) flows from the 1  1 state to-
does not yield convergence with the mass balance equation, and ward the w–w state which carries with it the properties of
hence, the results produced therein deviate from those obtained the ambient, i.e., the 1-state.
through the corrected energy balance equation used in this paper. (2) a mass flow rate of (gAo þ m_ w ) crossing the 1  1 state
Therefore, this paper aims at presenting the modified Reynolds away from the w–w state which carries with it the proper-
flow model and thereby achieving a correct prediction of system ties of the w-state.
behavior under varying conditions. It should be noted that the
symbols and notations used in this paper are intentionally kept where Ao is the curved surface area 2pro H of the external pot
similar to those used in Ref. [5] unless a different notation seemed from where evaporation occurs. Under these assumptions, all phe-
must. This is the done for the sake of the readers to make an easy nomenon of the real flow in the considered phase and their effects
comparison of the two models. on the heat and mass transfer processes at the w–w surface will
remain unaffected by these hypotheses.
In addition to the above hypotheses which were primarily used by
2 Reynolds Flow Model Date [5], another hypothesis is required to accommodate for the con-
vective heat transfer. The hypotheses states that
The Reynolds flow model is a mathematical approach for
describing the convective mass transfer processes across an
(3) the mass flow rate of gAo encompasses the convective
interface by developing relations that are related to real transport
effects of neighboring phase.
phenomenon with reference to the mass transfer coefficient (g)
[6]. In the present case, the interface is formed primarily between It must be noted that the inclusion of the term gAo itself encom-
the transferred substance state (water) and the considered phase passes the heat due to natural convection in Fig. 2 which was
(stagnant surrounding air). The outer surface of the outer pot, referred to as Q_ nc in Ref. [5]. Moreover, inclusion of this hypothe-
which is considered wet at all times, is designated by w–w state. sis leads to convergence of iteratively solved equations to arrive
Considered phase is designated as 1  1 state to indicate that at a unique solution for evaporation rate of water. In Ref. [5], the
the surrounding is infinite. This state is imaginary as it is not convergence between the evaporation rate obtained from mass
physically present in ambient surroundings. The major heat and and energy conservation principle, respectively, could not be
mass transfer occurs at the interface of these two states. The hot achieved due to exclusion of this hypothesis while formulating the
and less humid air from the 1  1 state flows toward the wet energy balance equation. One of the major contributions in this
pot surface, water evaporates into the air resulting in cooling the paper is to include this hypothesis for achieving convergence and
pot surface. Figure 2 shows the different states that are present in accurate predictions of the results.

031006-2 / Vol. 8, SEPTEMBER 2016 Transactions of the ASME

Downloaded From: http://thermalscienceapplication.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jtsebv/935128/ on 03/05/2017 Terms of Use: http://www.as


2.1 Determination of Evaporation Rate From Mass Con- where B* is a dimensionless quantity determined by the following
servation Principle. The mass conservation principle implies that expression:
the evaporation of water in the form of vapor is balanced between " #
the transferred phase and the considered phase and that no mass is  0:622ð1 þ 0:05 H Þ þ ð 0:0189 H  0:622Þxv;w
generated through this process. The derivation for mass conserva- B ¼ (14)
Bim ð0:622 þ 0:378 xv;w Þ
tion shown below can also be referred in Date [5]. To present the
model in totality and for clear understanding of the model, we
have included the derivation using simple notations and necessary In Eq. (14), Bim  g=ðqw K H Þ may be viewed as mass transfer
details. Thus, invoking the mass conservation law between the Biot number. It should be noted that Eq. (13) is different than Eq.
w–w state and 1  1 state, we get (4) mainly due to the inclusion of Darcy’s resistance in Eq. (13)
which better depicts the mass transfer in porous media. The
gAo xv;1 þ m_ w;m ¼ ðm_ w;m þ gAo Þxv;w (3) authors would like to point out that B* used in this paper differs
from the one used in Ref. [5].
Upon rearranging the equation, we get
2.2 Determination of Evaporation Rate From Energy
m_ w;m ¼ gAo  B (4) Conservation Principle. Equating the energy transfer confined
by the control volume bounded by w–w and 1  1 states, we
where can write the energy balance as
xv;w  xv;1
B¼ (5) gAo h1 þ m_ w;e hMw ¼ Q_ rad þ ðm_ w;e þ gAo Þhw (15)
1  xv;w

Since, the Lewis Number (Le) 1 for air–water vapor system [1], where gAo h1 is the heat going inside the considered phase from
the mass transfer coefficient “g” (kg m2 s1) can be calculated the imaginary 1  1 state. This heat energy is the actual heat
from natural convection heat transfer coefficient “ao” from the transferred due to natural convection.
correlation given in Ref. [7] m_ w;e hMw is the heat going out of the transferred phase.
Q_ rad is the heat due to radiation that interacts with the w–w
ao state.
gffi (6)
Cpm ðm_ w;e þ gAo Þhw is the total heat out due to evaporation of water
into the considered phase.
Cpm ¼ Cpa ð1  xv;mean Þ þ Cpv xv;mean (7) Once we obtain a converged solution for evaporation rate of
water (m_ w ) using Eqs. (13) and (15), we use the unique solution
xv;mean ¼ 0:5ðxv;1 þ xv;w Þ (8) for evaporation rate for estimating the other unknown variables
(refer Fig. 3). Therefore, on considering the control volume
In Eq. (4), B is known as the Spalding number. between M–M and 1  1 state and writing the energy balance
Further, this evaporation rate meets with the resistance offered equation, we get
by the outer clay pot wall thickness. Therefore, using Darcy’s
law, the average mass flow rate over height H is gAo h1 þ m_ w hM þ Q_ load ¼ Q_ L þ ðm_ w þ gAo Þhw (16)
   where m_ w (kg/s) is the converged evaporation rate of water from
qw kp Ao pa p H=2 pw;sat
m_ 0w;m ¼  (9) mass and energy conservation equations.
lw bo pa pa
hMw ¼ Cpw ðTw  Tref Þ (17)
where kp (m2) is permeability. Note that pH/2/pa ¼ (1 þ 0.05 H)
because pa ¼ 10 m of water. Also, since saturation conditions pre- h1 ¼ Cpa ðT1  Tref Þ þ ½ðCpv  Cpa ÞðT1  Tref Þ þ kref xv;1
vail at the w–w surface, pw;sat =pa ¼ xv;w =ð0:622 þ 0:378xv;w Þ.
Equation (9) can, therefore, be written as (18)
" #
0:622ð1 þ 0:05 H Þ þ ð 0:0189 H  0:622Þxv;w
m_ 0w;m ¼ qw K H Ao
0:622 þ 0:378xv;w
(10)

where K H ¼ ðkp pa Þ=ðlw bo Þ is hydraulic conductivity expressed in


(m s1).
Expressing Eq. (10) in the form of Eq. (4), we get

q KH
m_ 0w;m ¼ gAo  w
g
" #
0:622ð1 þ 0:05 H Þ þ ð 0:0189 H  0:622Þxv;w

0:622 þ 0:378xv;w
(11)
" #
0:622ð1 þ 0:05 H Þ þ ð 0:0189 H  0:622Þxv;w
m_ 0w;m ¼ gAo 
Bim ð0:622 þ 0:378 xv;w Þ
(12)

m_ 0w;m ¼ gAo  B (13) Fig. 3 Solution procedure for obtaining the temperature

Journal of Thermal Science and Engineering Applications SEPTEMBER 2016, Vol. 8 / 031006-3

Downloaded From: http://thermalscienceapplication.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jtsebv/935128/ on 03/05/2017 Terms of Use: http://www.as


hw ¼ Cpa ðTw  Tref Þ þ ½ðCpv  Cpa ÞðTw  Tref Þ þ kref xv;w (19) Adding last two Eqs. (32) and (33) will give Tcold in terms of Tw:
 
Q_ load 1 bi
hM ¼ Cpw ðTM  Tref Þ (20) Tcold  Tw ¼ þ
Ai ai kclw
and Q_ L is the heat conduction in the clay walls and sand–water  
bo ri lnðro =ri Þ  _
layer. Combining Eqs. (15) and (16) gives þ þ Q load  Q_ L (34)
kclw Ao ksw Ai
Q_ load  Q_ L ¼ m_ w ðhMw  hM Þ  Q_ rad (21)
From arranging Eqs. (29) and (32), we get
     
Q_ load  Q_ L ¼ m_ w Cpw Tw 
Ti þ To
 Q_ rad Ti þ To 1 2bo ri lnðro =ri Þ  _
2
(22) Tw  ¼ þ Q load  Q_ L (35)
2 2 kclw Ao ksw Ai
Radiative heat transfer is given by the following relation: Substituting the above Eq. (35) in Eq. (22), we get
 Q_ rad
Q_ rad ¼ e r AðT1
4
 Tw4 Þ (23) Q_ load  Q_ L ¼     (36)
m_ w Cpw 2bo ri lnðro =ri Þ
1þ þ
where emissivity, e ¼ 1, Stefan Boltzmann constant (r) ¼ 5.67 2 kclw Ao ksw Ai
 108 W m2 K4
Finally, on substituting Eq. (36) in Eq. (34), we get the cold zone
temperature with respect to evaporation rate and temperature of
Q_ load ¼ mfood  q_ load (24)
water at steady-state
 
where q_ load is the respiratory heat load of the food item stored and bo ri lnðro =ri Þ 
mfood is the mass of food in kg. The external heat transfer coeffi- Tcold  Tw ¼  þ  Q_ rad
kclw Ao ksw Ai
cient “ao ” for calculating the mass transfer coefficient “g” can be   1
found out using the Nusselt number, NuH. m_ w Cpw 2bo ri lnðro =ri Þ
 1þ þ
2 kclw Ao ksw Ai
9:81bðT1  Tw Þ H 3 1 _  
Gr H ¼ ; b¼ ; Q 1 bi
t2a Tmean (25) þ load þ (37)
Ai ai kclw
Tmean ¼ 0:5ðTw þ T1 Þ
Since evaporation can achieve a minimum temperature equiva-
Nusselt’s number for natural convection from a vertical plate can lent to the wet bulb temperature, the efficiency of a passive cool-
be estimated by [8] ing system is given by the relation given in Eq. (1).

Nu H ¼ 0:59ðGr H PrÞ0:25 (26) 2.3 Solution Procedure. The above equations for evapora-
tion rates are solved iteratively in order to achieve at a concord-
ao H ance. Rest of the parameters are either determined through
Nu H ¼ (27) experimental evidence or deduced through specific methods
ka
Table 1 Values for parametric estimation
The heat transfer coefficient “ao ” obtained from Eq. (27) is substi-
tuted in Eq. (6) to obtain “g.” The heat from conduction in the wet Value for the
clay and sand layers can be found out using the heat conduction Parameter Value reference case
equation as given below:
  Thermal conductivity of clay 0.38–2.5a,b 1.5
ksw Ai (kcl) (W m1 K1)
Q_ L  Q_ load ¼ ðTo  Ti Þ (28)
ri lnðro =ri Þ Thermal conductivity of sand 0.20–0.32c 0.26
  (ks) (W m1 K1)
kclw Ao Thermal conductivity of water 0.6096 0.6096
Q_ L  Q_ load ¼ ðTw  To Þ (29) (kw) (W m1 K1)
bo
Hydraulic conductivity 2  109–3  108a 8  109
(KH) (m s1)
where effective thermal conductivity for two-phase media of solid Void fraction for clay (hcl) 0.2–0.64 0.41
and fluid is given by [9] Void fraction for fine sand (hs) 0.29–0.46 0.37
Outer pot radius (ro) (m) 0.1–0.3 0.15
ksw ¼ ks1hs  kwhs (30) Outer pot thickness (bo) (m) 0.004–0.012 0.005
Inner pot thickness (bi) (m) 0.004–0.012 0.005
1hcl Pot height (H) (m) 0.2–1.0 0.3
kclw ¼ kcl  kwhcl (31) Specific heat of air (Cpa) 1005 1005
(J kg1 K1)
Specific heat of water (Cpw) 4186 4186
Combining Eqs. (28) and (29), we get (J kg1 K1)
  Specific heat of water vapor 1880 1880
bo ri lnðro =ri Þ  _
(Cpv) (J kg1 K1)
Tw  Ti ¼ þ Q L  Q_ load (32)
kclw Ao ksw Ai Latent heat of water vapor 2503 2503
(kref) (kJ kg1)
Heat transfer across S–S surface will give the cold zone tempera- Total heat load ðQ_ load Þ (W) 0.4–3 1
ture according to the following relation: Ambient temperature (T1) (  C) 30–50 40
RH (%) 10–80 10
 
Q_ load 1 bi a
Reference [5].
Tcold  Ti ¼ þ (33) b
Reference [10].
Ai ai kclw c
Reference [11].

031006-4 / Vol. 8, SEPTEMBER 2016 Transactions of the ASME

Downloaded From: http://thermalscienceapplication.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jtsebv/935128/ on 03/05/2017 Terms of Use: http://www.as


_ load 51 W, Tdp 5 2.58  C
Table 2 Results obtained for the reference case: Tw 5 40  C, RH 5 10%, Q

Case 1 Tw ¼ 18.2008 m_ w ¼ 2.2685  106 ao ¼ 4.4459 g ¼ 0.0044


To ¼ 18.2003 Q_ L ¼ 1.0392 Q_ rad ¼ 0.0393 B ¼ 0.0089
Ti ¼ 18.1841 Q_ in ¼ 64.1674 Bim ¼ 548.7287 B* ¼ 0.0018
Tcold ¼ 19.7977 COP ¼ 0.0156 gth ¼ 0.5398 xv,w ¼ 0.0132
Case 2 Tw ¼ 18.2008 m_ w ¼ 2.2685  106 ao ¼ 4.4459 g ¼ 0.0044
To ¼ 18.2001 Q_ L ¼ 1.0389 Q_ rad ¼ 0.0393 B ¼ 0.0089
Ti ¼ 18.1084 Q_ in ¼ 64.1674 Bim ¼ 548.7287 B* ¼ 0.0018
Tcold ¼ 19.7323 COP ¼ 0.0156 gth ¼ 0.5416 xv,w ¼ 0.0132

understood from relevant literatures. The solution procedure Q_ load ¼ 1 W: (1) with all the parameters identical as were consid-
includes the steps as given in Fig. 3. ered in Ref. [5] and (2) with parameters related to effective ther-
mal conductivities of clay þ water and sand þ water were revised
3 Results and Discussions for better accuracy in computation. The void fraction for clay
Several parameters governing the performance of the system samples usually vary in the range of 0.2–0.64 [12] and that for
are enlisted in Table 1. The values for some parameters have been sand varies in the range of 0.29–0.46 [13]. For the reference case,
obtained from the literature while others have been varied pur- void fraction of 0.41 for clay and 0.37 for sand and thermal con-
posefully to study their effect on the system. ductivity of water [14], i.e., equal to 0.6096 W m1 K1, were
considered for estimation of cold zone temperatures at various
3.1 Reference Case. Table 2 shows the results obtained for ambient conditions. The results have been obtained by iterative
the two reference cases at T1 ¼ 40  C, RH ¼ 10%, and determination of Tw. The trend for Ti, To, and Tw was found to be

Table 3 Effect of T‘ and RH at Q_ load 5 1 W

RH (%) m_ w  106 (kg s1) Tw ( C) a0 (W m2 K1) B* Tcold ( C) Tdp ( C) gth COP

T1 ¼ 30  C
10 2.2781 13.0247 4.2030 0.0019 14.71 4.95 0.4375 0.0231
20 2.2700 15.4083 4.0429 0.0020 17.16 4.57 0.5049 0.0205
30 2.2619 17.6319 3.8756 0.0021 19.46 10.51 0.5407 0.0185
40 2.2536 19.7123 3.6980 0.0022 21.63 14.91 0.5543 0.0172
50 2.2453 21.6622 3.5058 0.0023 23.69 18.42 0.5450 0.0163
60 2.2369 23.4926 3.2927 0.0024 25.65 21.37 0.5035 0.0157
70 2.2285 25.2134 3.0471 0.0026 27.55 23.92 0.4022 0.0155
80 2.2199 26.8340 2.7461 0.0029 29.43 26.16 0.1479 0.0159
T1 ¼ 35  C
10 2.2737 15.6168 4.3309 0.0019 17.23 1.18 0.4912 0.0188
20 2.2629 18.5132 4.1541 0.0020 20.20 8.66 0.5620 0.0164
30 2.2519 21.1676 3.9713 0.0020 22.93 14.81 0.5976 0.0147
40 2.2408 23.6072 3.7795 0.0021 25.47 19.35 0.6093 0.0136
50 2.2296 25.8551 3.5741 0.0023 27.83 22.99 0.5975 0.0128
60 2.2183 27.9320 3.3483 0.0024 30.04 26.05 0.5538 0.0123
70 2.2068 29.8570 3.0900 0.0026 32.15 28.69 0.4511 0.0121
80 2.1953 31.6470 2.7746 0.0029 34.21 31.02 0.1982 0.0124
T1 ¼ 40  C
10 2.2685 18.2008 4.4459 0.0018 19.73 2.58 0.5417 0.0156
20 2.2541 21.6695 4.2513 0.0019 23.28 12.74 0.6133 0.0134
30 2.2396 24.7767 4.0532 0.0020 26.47 19.09 0.6471 0.0119
40 2.2247 27.5727 3.8483 0.0021 29.37 23.79 0.6558 0.0109
50 2.2098 30.1012 3.6318 0.0022 32.02 27.56 0.6415 0.0102
60 2.1947 32.4009 3.3964 0.0024 34.46 30.72 0.5970 0.0098
70 2.1794 34.5050 3.1293 0.0025 36.75 33.46 0.4969 0.0096
80 2.1641 36.4415 2.8050 0.0028 38.97 35.87 0.2494 0.0098
T1 ¼ 45  C
10 2.2624 20.7934 4.5496 0.0018 22.23 6.31 0.5885 0.0131
20 2.2434 24.8805 4.3368 0.0019 26.40 16.79 0.6593 0.0110
30 2.2242 28.4447 4.1245 0.0019 30.06 23.36 0.6904 0.0097
40 2.2047 31.5801 3.9087 0.0021 33.31 28.22 0.6967 0.0088
50 2.1849 34.3654 3.6837 0.0022 36.22 32.12 0.6817 0.0082
60 2.1650 36.8646 3.4417 0.0023 38.87 35.39 0.6379 0.0078
70 2.1450 39.1288 3.1693 0.0025 41.33 38.22 0.5413 0.0077
80 2.1248 41.1978 2.8408 0.0028 43.67 40.72 0.3107 0.0078
T1 ¼ 50  C
10 2.2550 23.4057 4.6435 0.0017 24.74 10.03 0.6320 0.0111
20 2.2303 28.1410 4.4130 0.0018 29.56 20.84 0.7010 0.0092
30 2.2052 32.1518 4.1882 0.0019 33.68 27.61 0.7289 0.0080
40 2.1798 35.6045 3.9636 0.0020 37.25 32.64 0.7344 0.0072
50 2.1540 38.6256 3.7325 0.0021 40.40 36.67 0.7202 0.0066
60 2.1281 41.3092 3.4859 0.0023 43.24 40.06 0.6791 0.0063
70 2.1019 43.7241 3.2104 0.0024 45.86 42.98 0.5897 0.0061
80 2.0756 45.9213 2.8801 0.0027 48.34 45.57 0.3747 0.0062

Journal of Thermal Science and Engineering Applications SEPTEMBER 2016, Vol. 8 / 031006-5

Downloaded From: http://thermalscienceapplication.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jtsebv/935128/ on 03/05/2017 Terms of Use: http://www.as


Fig. 4 (a) Estimated variation in gth with RH (%). (b) Estimated value of DT (T‘ 2 Tcold) with RH (%).

same as was predicted in Ref. [5] with Ti < To < Tw. Dew point wet sand as 2 W m1 K1. Reference case 2 makes use of appro-
temperature corresponding to these ambient conditions was found priate correlations for effective thermal conductivities of clay and
to be lower than Tcold as expected, thereby yielding an efficiency sand (Eqs. (30) and (31)), taking into account their respective
of 53.98% for reference case 1 and 54.16% for reference case 2. porosity values. This also provides higher system efficiency in
The driving force for mass transfer (B) is reduced to B* ¼ 0.0018 comparison to reference case 1.
in both the cases due to Darcy’s resistance. It must be noted that
although the values of different parameters for case 1 and 2 are 3.2 Effect of Ambient Temperature and Humidity. Table 3
similar, it was necessary to use the case 2 as the actual reference shows the data obtained through iterative solving of heat and mass
case for this study. This is mainly because case 1 does not account balance equations to predict the cold zone temperature (Tcold) and
for the wet clay pot which is present in practical systems. It only effectiveness of the system (gth). It should be noted that the values
accounts for wet sand and uses effective thermal conductivity of presented in Table 3 are computed based on the reference case 2.

Fig. 5 Variation of DT and gth with respect to (a) outer pot radius, (b) pot thickness, and (c)
pot height

031006-6 / Vol. 8, SEPTEMBER 2016 Transactions of the ASME

Downloaded From: http://thermalscienceapplication.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jtsebv/935128/ on 03/05/2017 Terms of Use: http://www.as


increase in humidity and temperature. This is due to the
humid air that has high mass fraction of water vapor in it and
has, therefore, less water uptake capacity to achieve saturation
without condensing.

3.3 Effect of Geometrical Parameters of the Pot-in-Pot


System. Based on computation carried out for ambient condi-
tions similar to the reference case 2, Figs. 5(a)–5(c) show the
effect of geometrical parameters such as clay pot thickness,
height and pot radius on the steady temperature difference
between the cold zone and the ambient (DT), and the system
efficiency (gth). It is to be noted that the pot radius was varied in
order to keep the thickness of the sand þ water layer constant,
i.e., (ro–ri) ¼ constant ¼ 0.075 m. For the convenience of repre-
sentation, in Fig. 5(a), only outer pot radius is mentioned in the
x-axis. It can be noted that with the increase in pot radius, the
steady-state temperature difference increases for smaller radii
_ load
Fig. 6 Typical variation in DT and gth with respect to Q and then gradually settles around a radius of 0.5 m. This obser-
vation is extremely useful in designing novel systems based on
It can be seen that in all cases, the cold zone temperature (Tcold) is evaporative cooling principles.
found to be higher than the dew point (Tdp) and it increases with Figure 5(b) shows that the thickness of the pot does not signifi-
increase in RH. Equation for determining the dew point tempera- cantly alter the temperature difference. With increase in thickness
ture is given in appendix. From the equation, it can be understood from 0.004 m to 0.012 m, the change in DT is only about 0.2  C.
that the dew point is a function of ambient temperature and RH. However, it is only valid to infer that the increase in thickness
Therefore, although the cooling temperature is minimum in case will cause greater thermal and mass transfer resistance for the
of lowest RH condition (i.e., 10%), the efficiency of the system is heat and water to flow across the layers of the pot. Thus, the total
not the maximum. Instead, it is found to be maximum at a RH of heat load may possibly increase the cold zone temperature despite
40% for the entire ambient temperature range. This is due to the evaporation occurring at the surface of the outer pot.
psychrometric behavior of air–water vapor system. It is known Figure 5(c) shows the effect of pot height on the system. It can
that water evaporates at all temperatures. However, only the ambi- be observed that the DT increases at a faster rate with initial
ent conditions determine the equilibrium temperature of water at increase in height from 0.2 m to 0.4 m and then increases gradu-
which cooling occurs. Tcold in all cases is greater than Tw due to ally beyond 0.4 m. This observation indicates that height does
presence of respiratory heat load in the inner pot. play an important role in achieving better performance. Increase
Figure 4(a) shows how the system efficiency varies with in height increases the effective area for evaporation of water
ambient conditions. It can be seen that efficiency is maximum which leads to higher temperature differences.
at RH of 40% for all the cases similar to the one presented in
Ref. [5]. However, the efficiency predicted with the modified 3.4 Effect of Heat Load. Respiratory heat load (q_ load ) varies
model is much higher than the one predicted in Ref. [5]. It is depending on the type of food item. Typical variation has been
worth mentioning here that strangely the COP values in case found to be in the range of 0.2–2 W kg1 of food [15]. Thus,
of Date’s model were found to increase with increasing RH at the total heat load (Q_ load ) was estimated as the sum of respira-
any particular temperature. This is practically not possible tory heat load and the heat leakage from the top and the bottom.
considering that the air enthalpy increases with increase in Figure 6 shows the effect of total heat load on the system per-
temperature and water vapor mass fraction. Therefore, the Q_ in formance. It can be seen as expected, both DT and gth decrease
will always increase with increase in RH and temperature, with increase in heat load for the ambient conditions similar to
and hence, the COP decreases. In Table 3, we present the the reference case 2. While increase in heat load could mean
COP values with expected trends obtained with the modified increase in the amount of food items stored or increase in the
model. Figure 4(b) plots the temperature difference between heat leakage from the top or bottom faces, the effect in either
cold zone and ambient versus the RH. The graph follows an case can only lead to increase in the cold zone temperature,
expected pattern, wherein DT decreases nonlinearly with thereby reducing the system efficiency.

Fig. 7 Variation in DT and gth with respect to (a) effective thermal conductivity of wet clay and
(b) effective thermal conductivity of wet sand

Journal of Thermal Science and Engineering Applications SEPTEMBER 2016, Vol. 8 / 031006-7

Downloaded From: http://thermalscienceapplication.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jtsebv/935128/ on 03/05/2017 Terms of Use: http://www.as


system efficiencies with and without Q_ rad is too high to be insig-
nificant. Therefore, it is concluded that Q_ rad is necessary for accu-
rate estimation of cold zone temperature.

4 Conclusions
The presented Reynolds flow model makes use of an energy
balance equation, wherein the mass flux due to enthalpy differ-
ence is considered to provide the convective heat transfer. This
is an important point to note as without this hypothesis, the heat
and mass balance equation would not converge for a single evap-
oration rate at any physically possible temperature of water.
Based on the results, it can be concluded that the pot thickness,
thermal, and hydraulic conductivities do not contribute signifi-
cantly toward cooling. Pot height, pot radius, heat load, and radi-
Fig. 8 Variation in DT and gth with respect to KH ation heat transfer, on the other hand, can significantly affect the
cold zone temperature. The temperature difference can reach a
maximum of close to 25  C for ambient conditions of 50  C and
3.5 Effect of Thermal Conductivities of Wet Clay and 10% RH, while minimum temperature difference could be seen
Sand. The effective thermal conductivity of wet clay based on close to 0.5  C for ambient conditions of 30  C and 80% RH.
Eq. (31) has been found to be in the range of 0.4613–1.4017 W Pot-in-pot system is an extremely befitting solution for problems
m1 K1. Similarly, from Eq. (30), the effective thermal conduc- arising due to lack of cold storage in unelectrified households in
tivity of wet sand has been found to lie in the range of rural areas. Based on these results, current work is in progress
0.3021–0.4062 W m1 K1. Under the reference case 2 condi- to develop a passive evaporative cooler using the pot-in-pot
tions, it can be seen from Fig. 7 that the effective thermal conduc- concept with built-in dehumidification unit that can achieve
tivity of wet clay as well as wet sand does not produce any maximum cooling capacity for any ambient condition. The
significant change in the DT. Therefore, the system efficiency in experimental and modeling results obtained based on this work
both the cases was found to be more or less constant. This is would be presented in future.
mainly due to low thermal conductivities of both clay as well as
sand that the temperature difference is only marginally affected.
Acknowledgment
3.6 Effect of Hydraulic Conductivity. The trend for the The authors would like to thank the Director, CSIR-SERC,
effect of hydraulic conductivity is depicted in Fig. 8. It can be Director, CSIR-CEERI, and the Scientist-in-Charge, CSIR-
seen that the hydraulic conductivity which, in turn, is dependent CEERI, Chennai, for their valuable support. We owe our sincere
on permeability, pot thickness, and viscosity of water does not gratitude to Professor Anil W. Date, Mechanical Engineering
provide any significant change in the DT. This is because the Department, IIT-Bombay, for providing useful inputs for this
hydraulic conductivity is directly proportional to the permeability study. We thank Mr. Meraj Khan and Mr. N. Balakrishna, Ph.D.
and inversely to the pot thickness. Thus, increase in hydraulic scholars, AcSIR, for helping in this study. Finally, we also thank
conductivity suggests that either the material has higher perme- Dr. Alok Paul, Senior Scientist, CSIR-CECRI, Chennai, for his
ability or has lesser thickness. Both of these may lead to increase valuable inputs in this study.
in heat entering the system from the ambient and thus reducing
the temperature difference. Nomenclature
A¼ surface area (m2)
3.7 Effect of Radiation Heat Transfer. Radiation heat trans- b¼ clay wall thickness (m)
fer is normally significant at higher temperature differences. B¼ Spalding number
Hence, it is necessary to understand whether or not the radiative Bim ¼ Biot number
heat transfer has a significant role in pot-in-pot system perform- Cp ¼ specific heat (J/kg K)
ance. Figure 9 shows the comparison of DT and system efficiency g¼ mass transfer coefficient (kg/m2 s)
for the reference case at 40  C ambient with and without incorpo- h¼ enthalpy of the mixture (J/kg)
ration of radiation effects (Q_ rad ) at varying percentage RH. The H¼ pot height (m)
graph clearly indicates that there is a significant shift both in DT k¼ thermal conductivity (W/m K)
as well as efficiency. In addition, while the peak efficiency with kp ¼ permeability (m2)
incorporation of Q_ rad is obtained at 40% RH, the same is obtained KH ¼ hydraulic conductivity (m/s)
at 70% RH in case of without Q_ rad . The difference between Le ¼ Lewis number
m¼ mass (kg)
mw ¼ evaporation rate (kg/s)
P¼ pressure (N/m2)
Q¼ heat transfer (W)
RH ¼ relative humidity (%)
T¼ temperature (  C)

Greek Symbols
a¼ heat transfer coefficient (W/m2 K)
b¼ volumetric coefficient (K1)
e¼ emissivity
k¼ latent heat (J/kg)
h¼ void fraction
_ rad
Fig. 9 Variation in DT and gth with and without Q gth ¼ thermodynamic efficiency

031006-8 / Vol. 8, SEPTEMBER 2016 Transactions of the ASME

Downloaded From: http://thermalscienceapplication.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jtsebv/935128/ on 03/05/2017 Terms of Use: http://www.as


 
l ¼ viscosity W1 pv;1
x ¼ mass fraction xv;1 ¼ ; where W1 ¼ 0:622 
1 þ W1 ptot  pv;1
 
Suffixes pv;1 RH psat F T1
¼ and ¼ exp ; where s ¼ 1 
a¼ air psat 100 pcr 1s Tcr
cl ¼ clay
clw ¼ clay þ water F ¼ a1  s þ a2  s1:5 þ a3  s3 þ a4  s3:5 þ a5  s4 þ a6  s7:5
cold ¼ inner pot environment
dp ¼ dew point
e¼ energy conservation principle a1 ¼ 7:85951783; a2 ¼ 1:84408295; a3 ¼ 11:7866497;
i¼ inner pot a4 ¼ 22:6807411; a5 ¼ 15:9618719; a6 ¼ 1:80122502
in ¼ input
L¼ conduction in sand þ water
m¼ mass conservation principle or mean
M¼ transferred substance state References
o¼ outer pot [1] Crawford, M., and Kays, W. M., 1993, Convective Heat and Mass Transfer,
rad ¼ radiation McGraw-Hill International Edition, New York.
[2] Mittal, A., Kataria, T., Das, G. K., and Chatterjee, S. G., 2006, “Evaporative
ref ¼ reference Cooling of Water in a Small Vessel Under Varying Ambient Humidity,” Int. J.
s¼ sand Green Energy, 3(4), pp. 347–368.
sat ¼ saturation [3] Camargo, J. R., Ebinuma, C. D., and Cardoso, S., 2003, “A Mathematical
sw ¼ sand þ water Model for Direct Evaporative Cooling Air Conditioning System,” Eng. Term.,
4, pp. 30–34.
v¼ vapor [4] Van der Sman, R. G. M., 2003, “Simple Model for Estimating Heat and Mass
w¼ water or w–w state Transfer in Regular-Shaped Foods,” J. Food Eng., 60(4), pp. 383–390.
1¼ ambient condition [5] Date, A. W., 2012, “Heat and Mass Transfer Analysis of Clay Pot Refriger-
ator,” Int. J. Heat Mass Transfer, 55(15–16), pp. 3977–3983.
[6] Spalding, D. B., 1963, Convective Mass Transfer, Edward Arnold, London.
Appendix [7] Arora, C. P., 2006, Refrigeration and Air Conditioning, Tata McGraw-Hill,
New Delhi, India.
(1) Dew point temperature can be found from Ref. [16]: [8] McAdams, W. H., 1954, Heat Transmission, 3rd ed., McGraw-Hill, New
York.
  [9] Woodside, W., and Messmer, J. H., 1961, “Thermal Conductivity of Porous
237:7  c 17:271  T1 %RH
Tdp ¼ where; c ¼ þ ln Media. I. Unconsolidated Sands,” J. Appl. Phys., 32(9), pp. 1688–1698.
17:271  c 237:7 þ T1 100 [10] Etuk, S. E., Akpabio, I. O., and Udoh, E. M., 2003, “Comparison of the Ther-
mal Properties of Clay Samples as Potential Walling Material for Naturally
(2) xv;w can be estimated for known Tw by the following Cooled Building Design,” J. Environ. Sci., 15, pp. 65–68.
equation [1]: [11] Smits, K. M., Sakaki, T., Limsuwat, A., and Illangasekare, T. H., 2009,
“Determination of the Thermal Conductivity of Sands Under Varying Moisture,
Drainage/Wetting, and Porosity Conditions—Applications in Near-Surface Soil
xv;w
3:416  106 þ ð2:7308  104 ÞTw Moisture Distribution Analysis,” Hydrol. Days, pp. 57–65.
[12] Hough, B., 1969, Basic Soil Engineering, Ronald Press Company, New York.
þ ð1:372  105 ÞTw2 þ ð8:2516  108 ÞTw3 [13] Das, B., 2008, Advanced Soil Mechanics, Taylor and Francis, London.
[14] Ramires, M. L. V., Nieto de Castro, C. A., Nagasaka, Y., Nagashima, A.,
ð6:9092  109 ÞTw4 þ ð3:5313  1010 ÞTw5 Assael, M. J., and Wakeham, W. A., 1995, “Standard Reference Data for the
Thermal Conductivity of Water,” J. Phys. Chem. Ref. Data, 24(3),
 ð3:7037  1012 ÞTw6 þ ð6:1923  1015 ÞTw7 pp. 1377–1381.
þ ð9:9349  1017 ÞTw8 [15] Cengel, Y. A., and Ghajjar, A. J., 2006, Heat and Mass Transfer: Fundamentals
and Applications, 4th ed., McGraw-Hill Ryerson, Toronto, ON, Canada.
[16] Wagner, W., and Pruss, A., 1993, “International Equations for the Saturation
(3) xv;1 can be estimated from known RH using relation given Properties of Ordinary Water Substance. Revised According to the International
by [16] Temperature Scale of 1990,” J. Phys. Chem. Ref. Data, 22(3), pp. 783–787.

Journal of Thermal Science and Engineering Applications SEPTEMBER 2016, Vol. 8 / 031006-9

Downloaded From: http://thermalscienceapplication.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jtsebv/935128/ on 03/05/2017 Terms of Use: http://www.as

You might also like