Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Applied Thermal Engineering 24 (2004) 2537–2555

www.elsevier.com/locate/apthermeng

Heat and mass transfer during adsorption of ammonia


in a cylindrical adsorbent bed: thermal performance study
of a combined parabolic solar collector, water heat pipe
and adsorber generator assembly
a,*
F. Aghbalou , A. Mimet c, F. Badia b, J. Illa a, A. El Bouardi d, J. Bougard e

a
Department of Computer and Industrial Engineering, University of Lleida, C/Jaume II, 69, 25001 Lleida, Spain
b
Energetic and Fluid Mechanics Laboratory, Sciences Faculty, B.P 2121, 39000 Tetouan, Morocco
c
Department of Environmental and Soil Sciences, University of Lleida, C/Jaume II, 69, 25001 Lleida, Spain
d
Thermal, Solar Energy and Environment Laboratory, Faculty of Sciences, Abdelmalek Essaadi University,
P.O. Box 2121, Tetuan 93000, Morocco
e
Solar Energy Research Centre, Polytechnic Faculty of Mons and Bruxelles University,
31, Boulevard Dolez, B-7000 Mons, Belgium
Received 18 December 2003; accepted 22 April 2004
Available online 28 July 2004

Abstract
In this paper we present the study of adsorption refrigerator which use an activated carbon-pair
ammonia. The ability of activated carbons to adsorb large mass of ammonia makes them ideal for use in
adsorption refrigeration and pump systems. These systems have not reasonable efficiency. In order to make
these systems economically viable, their size must be reduced. This implies a need for a rapid heating and
cooling the adsorbent/refrigerant pair. However, the main problems to be overcome is related to the poor
heat transfer in the adsorbent bed. So, it is necessary to study and understand the heat and mass transfer
within the bed and to improve it. A detailed model of heat and mass transfer into the generator has been
developed. For a given heat flux, temperature and adsorbed mass have been computed in every point at
each step time along the adsorbed bed (generator). Experimental installation simulating an adsorption
machine working within a temperature ranging from 20 to 250 C and pressure ranging from 0 to 2.5 · 106
Pa, allows for identification of the generator’s equivalent thermal conductivity and internal heat transfer
coefficient. These two parameters are then used to simulate thermal performance of a design whose features
include the insertion of stainless steel water heat pipe (HP’s) condensers into the generator. The HP’s

*
Corresponding author. Tel.: +34-973-702-744/764; fax: +34-973-702-702.
E-mail addresses: fouad@diei.udl.es (F. Aghbalou), mimet@fst.ac.ma (A. Mimet), fbadia@macs.udl.es (F. Badia),
jilla@macs.udl.es (J. Illa), abouardi@fst.ac.ma (A. El Bouardi), bougard@fpms.fpms.ac.be (J. Bougard).

1359-4311/$ - see front matter  2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.applthermaleng.2004.04.009
2538 F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555

evaporator heat input is of solar origin using a compound parabolic collector (CPC). Nominal Solar
coefficient of performance, COPs ¼ 14.37% obtained through both Adimensional Exergy Loss (AEL), and
COP study, shows the competitiveness of the proposed design.
 2004 Elsevier Ltd. All rights reserved.

Keywords: Solar energy; Refrigeration; Adsorption; Activated carbon-ammonia; Heat pipe

1. Introduction

Since environmental constraints have limited CFC and HCFC use, systems utilising heat energy
as a primary source such as adsorption refrigeration machine have emerged. Unlike absorption
systems, which use chemical reaction, the adsorption systems use physical phenomenon, which is
somewhat similar to the process of condensation. Equilibrium between the solid surface and the
gas molecules is usually rapidly attained and easily reversible because the energy requirements
are small. Solid adsorption cooling machine have been extensively studied [1]. In fact, they
constitute very attractive solutions for both industrial waste heat recovery and exploitation of
renewable energy such as solar energy. Adsorption cooling machines consist essentially of an
evaporator, condenser, valves, adsorbent fluid and a generator containing a porous medium, for
instance, AC/ammonia. An ideal cycle for a solar cooling machine can be described as follows
(Fig. 1).

• Phase A. When heating the reactor, the pressure increases whereas the adsorbed mass remains
constant. This phase is described as isosteric-heating.
• Phase B. The mixture pressure reaches the saturation pressure of the pure refrigerant at the
condenser’s temperature. The desorption begins and the pressure remains constant until the
adsorbent reaches its maximum temperature where the refrigerant condenses. It is the isobaric
condensation–desorption phase.
• Phase C. The reactor is cooled, and consequently the pressure decreases and follows an isoster.
It is the isosteric-cooling phase. the pressure of the reactor decreases until it reaches the pres-
sure of the evaporator.
• Phase D. The adsorption occurs inducing a cold production in the evaporator. It is the isobaric
evaporation–adsorption phase.

Intermittent adsorption cycles avoid certain problems associated with intermittent liquid
absorption systems, such as the need for a rectifier to strip the generated refrigerant of small
quantities of vaporised absorbent, and the need for check valves liquid seals to ensure mixing of
the absorbent during absorption. They represent the best system in terms of reliability. Against
this there are disadvantage of generator bulk and much poorer heat transfer compared with liquid
systems.
Critoph [2,3] has proposed two separate adsorption refrigeration systems operating out of
phase so that when one adsorber is being heated, the other cools to ambient, re-adsorbing its
refrigerant and producing useful cooling in the evaporator. However, the difficulties and disad-
vantages are that the heat transfer fluxes must be very high that the system must become complex,
F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555 2539

Nomenclature

a absorptance
T temperature, C
P pressure, bar
t time, min
L length, m
m mass, kg
c specific heat, KJ Kg1 K1
Q power, W
s entropy, kJ/kg
h heat transfer coefficient, W m2 K1
r, R radius, m
U internal energy, kJ
u specific internal energy, kJ kg 1
D diameter, m
hh specific enthalpy, kJ kg1
q flow rate, kg s1
S surface, m2
e thickness, m
E exergy, kJ
Hg global daily radiation, kW h m2
I irradiation, W m2
Subscripts
O outlet
I inlet
s solid
S saturation
g gas
ge generator
gen generation
gb global
gb, c global cold side
gb, h global hot side
gb, m global-mean
a adsorption (adsorbed, adsorbent)
B beam
D diffuse
d dead state
opt optimal
i internal
e external
2540 F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555

oil oil
ss stainless steel
eq equivalent
app apparent
hp heat pipe
u useful
re reflector
r,ði; jÞ radiative transfer between surfaces i and j
cv,ði; jÞ convective transfer between surfaces i and j
gl glass
v vapour
amb ambient
wick wick
ad adiabatic
ev evaporator (evaporation)
con condenser (condensation)
M, a maximal of adsorbent
m, a minimal of adsorbent
mean mean
0 initial
tot total
Greek symbols
k conductivity, W m1 K1
e porosity
DH isosteric heat, kJ kg1
q density, kg m3
a volume fraction of the adsorbed phase
U flux, W
r deviation
qo reflectance
s transmittance
g efficiency
Abbreviations
AEL adimensional exergy loss
HP heat pipe
COP coefficient of performance
COPs solar coefficient of performance
AC activated carbon
CPC compound parabolic collector
Cc concentration ratio
F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555 2541

Fig. 1. Ideal intermittent adsorption cycle.

requiring switching both the heat input and the cooling to ambient from one bed to another.
Modeling of such machines requires a good knowledge of a heat and mass transfer through the
porous bed. Several studies have been carried out. Among them we mention: Al Mers et al. [4]
have studied, through a mathematical model, the heat and mass transfer in a fixed bed of AC
reacting with Ammonia. They have used the Boubnov–Galerkin method. Sun et al. [5] have
determined the temperature and distribution within a spherical grain of the solid adsorbent under
a pressure pulse of adsorbable vapour. Kanolt [6] has carried out a theoretical and experimental
study of the non-isothermal adsorption of steam in a insulated zeolite grain.
Ruthven et al. [7] have studied the kinetics of a non-isothermal sorption and treated the heat
diffusion equation considering as uniform the grain temperature during the adsorption–desorp-
tion phenomena. Chihara et al. [8] have studied the generation effect on measurement of
adsorption rate. Guilleminot et al. [9] have treated the heat and mass transfer in a non-isothermal
fixed bed of AC/methanol. Wang et al. [10] have studied the performance of activated carbon–
methanol adsorption systems concerning heat and mass transfer. Results show that the heat
transfer coefficient of the solidified bed are much higher than that of granular bed. The design of
gas flow channels in adsorbers is very important to the performance of the mass transfer and the
whole system. Critoph et al. [11] have proposed a modified version of Zehener–Bauer model of
thermal conductivity of an AC bed exposed to adsorbable gas. In a later work [12], Critoph et.al.
suggested a conceptual model for a sorption refrigeration system using monolithic carbon and
ammonia. The model accounts for the heat and mass resistances and fluid heat transfer coeffi-
cients. It was found that there is a limit to the minimum heat transfer fluid channel thickness in
order to prevent the pumping power required from becoming too large relative to the cooling
power provided.
The aims of the present paper are:

• To contribute to a better understanding of heat and mass transfer into a porous medium,
constituted by AC/ammonia through a simple mathematical model. The generator’s thermal
2542 F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555

conductivity and the internal heat transfer coefficient are identified by comparing with experi-
mental results.
• To study numerically thermal performances, in terms of COPs and AEL, of the solar adsorp-
tion machine’s new design consisting in a CPC–HP–generator assembly. Thermal conductivity
and internal heat transfer coefficient are those identified experimentally.

2. Model

The considered cylindrical porous medium is constituted with grains of activated carbon
reacting by adsorption with ammonia, Fig. 2. The phases existing in the porous medium are: solid
phase constituted by carbon grains, gaseous phase, and adsorbed phase.

Fig. 2. Cylindrical double casing generator.


F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555 2543

The following assumptions have been adopted:

• The porous medium properties have a cylindrical symmetry.


• All phases are continuously in thermal, mechanical and chemical local equilibrium.
• The pressure is uniform.
• The heat transfer is radial and the convection heat transfer due to the radial mass transfer is
neglected.
• The conduction heat transfer in the medium can be characterised by an equivalent thermal con-
ductivity, keq .

All these assumptions have been validated by direct experimentation. The following data are
necessary for the model:

• Thermodynamics equilibrium properties of the reactive medium components. Activated carbon


[13], ammonia gas [14], adsorbed ammonia [15].
• Adsorption isotherms. The Dubinin equation [16] has been used.

3. Energy balance in the porous medium

The energy conservation equation applied for a layer of a porous medium (AC/ammonia) of a
thickness dr, Fig. 3, gives
dðdU Þ X X
þ ðqO hhO Þ  ðqI hhI Þ ¼ U ð1Þ
dt O I

The medium porosity, e, is defined as the relation between the available volume for the fluid
(gaseous and adsorbed phases) and the total volume of the medium.
The internal energy dU appearing in Eq. (1) may expressed as a function of the medium phases:
solid, adsorbed and gaseous, as
dU ¼ 2prdrL½ð1  eÞqs us þ ðe  aÞqg ug þ aqa ua  ð2Þ
where 1  e is the volume fraction of the solid phase.

Fig. 3. Experimental and computed temperatures into the generator. (keq ¼ 0:431 W m1 K1 , hi ¼ 33:45 W m2 K1 ).
2544 F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555

Eq. (1) is expressed as follows:


o
ð2prdrL½ð1  eÞqs us þ ðe  aÞqg ug þ aqa ua Þ  qðr þ dr; tÞhhg ðT ðr þ dr; P ÞÞ
ot
þ qðr; tÞhhg ðT ðr; P ÞÞ
 2 
o T oT
¼ keq 2prdrL þ ð3Þ
or2 ror
where q is the ammonia flow rate within the layer, r is the layer radius, dr is the layer thickness, P
is the reactor pressure, qapp is the adsorbent apparent density, qa is the adsorbed phase density,
defined by relation a ¼ ma qapp =qa and ma is the adsorbed mass in kg of ammonia per kg of
activated carbon.
The ammonia gas enthalpy is given by
hhg ðT Þ ¼ hha ðT Þ þ DHa ðT ; mÞ
where hha ðT Þ is the adsorbate enthalpy at temperature T and DHa ðT ; mÞ is the adsorption heat of
ammonia on activated carbon.

4. Mass conservation

Heat transfer induce a mass transfer in the porous medium. In fact, the hottest layer will desorb
the ammonia gas which will be adsorbed by the colder layers. This mass transfer will participate in
the heat transfer process, since the gas that will be adsorbed on the cold layer is hotter, and also
because of the exothermal character of adsorption. The difference between ammonia gas mass
entering and leaving the layer under consideration is equal to the adsorbed ammonia gas
remaining in the layer. The ammonia mass conservation equation is formulated as follows:
 
o oq
ð2prdrL½ðe  aÞqg þ aqa Þ ¼ qðr; tÞ þ qðr þ dr; tÞ ¼ dr ð4Þ
ot or

5. Combined heat and mass transfer equation

Energy equation (3) combined with the mass conservation equation yields:
oT o½ðe  aÞqg  P oðaqa Þ P o½aqa 
½ð1  eÞqs cs þ ðe  aÞqg cg þ aqa ca     DHa ðT ; P Þ
ot ot qg ot qa ot
 
o2 T oT
¼ keq þ ð5Þ
or2 ror
The five terms of Eq. (5) represent, respectively:

• gas, adsorbed and solid heating;


• gas elastic energy;
F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555 2545

• adsorbed elastic energy;


• adsorption energy;
• heat conduction.

Eq. (5) written for all the layers contained in the cylinder gives a non-linear system of partial
derivatives equation, which is completed by initial and boundary conditions.

5.1. Initial condition

At t ¼ 0, the temperature distribution in the porous medium is equal to T0 .


T ðr; oÞ ¼ T0 ðr ¼ 0; . . . ; RÞ.
T ðr; oÞ: temperature of layer of radius r at the time t ¼ 0.
T0 : initial temperature.
R: cylinder radius containing the porous medium.

5.2. Boundary conditions

Boundary conditions are


 
oT
for r ¼ 0 ðaxial symmetryÞ; ¼0
or r¼0
8  
> oTss
>
< mss css ¼ he Se ðToil  Tss Þ  hi Si ðTss  T Þ
ot  
for r ¼ R;
>
> oT
: hi ðTss  T Þ ¼ keq
or r¼R
where he is the external heat transfer coefficient between heating oil and metallic wall of the heat
exchanger; hi is the internal heat transfer coefficient between metallic wall and porous medium; T
is the porous medium temperature; Se is the external heat exchanger surface; and Si is the internal
heat exchanger surface.
The obtained non-linear set of equations is solved using an implicit finite difference scheme.

6. Experiment

6.1. Experimental device

The validity of the numerical results has been tested with a tubular generator constructed for
this purpose, Fig. 2. The cylindrical generator (Di ¼ 53 mm, L ¼ 250 mm) with a double stainless
steel envelope was heated by a thermal oil, which flowed along the 0.004 m thickness of the inter-
envelope space, with a flow rate of 1.5 l/min. The thermostat temperature ranged between 20 and
250 C. The generator cover contains two holes allowing the inlet or outlet of the ammonia gas.
The temperature was measured by means of six thermocouples in different points of the porous
medium. The respective positions on the cylinder radius and axis ri and zi , respectively, are given
2546 F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555

Table 1
Physical properties of the considered AC
Property Units Value
2 1
Specific area m g 1050–1150
Apparent density g cm3 0.48–0.54
Particle density g cm3 0.75–0.80
Real density g cm3 2.0–2.2
Pores volume g1 cm3 0.87–0.85
Specific heat at 100 C kcal kg1 C1 0.20–0.25

in Fig. 2. The generator was filled with 274 g of the activated carbon in particle form having a
mean diameter of 2 mm, into a volume of 548 cm3 . Physical properties of the AC are given in
Table 1. The void fraction was of about 71%. Pressure is measured at the two edges of the
generator.

6.2. Experimental result: temperature profiles

The generator is heated under vacuum and under pressure (closed and open). The temperatures
have been measured in the porous medium. The numerical solutions of the general energy
equation (5) have been compared with experimental results. The coefficients keq and hi have been
obtained minimising the objective function defined as the mean relative deviation r
j j
1 XM X 3
Tk;i  Texp;i
r¼ j ð6Þ
N j¼1 k¼1 Texp;i

j j
N is the number of comparison points, Tk;i and Texp;i are the computed and experimental tem-
perature respectively at position ri and at the instant j for the thermocouple number k. M is the
experiment duration.

6.2.1. Generator under vacuum


After introducing 274 g of activated carbon in the generator, we create a vacuum (1.33 Pa), in
order to desorb water vapour and other impurities in the carbon. Then we heat the reactor up to
150 C and we connect it to vacuum pump. When the vacuum is obtained, we begin to heat the
generator. Results are presented in Table 2.

Table 2
Results for a generator under vacuum
Temperature range (C) hi (W/m2 K) ke (W/m K) r (%)
½22; 39 25.65 0.276 1.2
½39; 55 26.35 0.298 0.8
½55; 72 26.85 0.319 1.4
½72; 91 28.75 0.335 1.2
F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555 2547

Table 3
Results for a closed generator
Temperature range (C) hi (W/m2 K) Pressure range (bars) ke (W/m K) r (%)
½25; 70 33.65 ½0:036; 0:33 0.348 1.5
½29; 70 34.35 ½0:582; 2:015 0.377 1.2
½30; 70 34.55 ½1:845; 6:074 0.386 1.8
½20; 70 33.45 ½2:97; 14 0.412 0.9

6.2.2. Closed generator reactor


By comparing experimental results with the computed ones for the closed generator under
pressure, we have obtained the coefficients keq and hi minimising the objective function r defined
above. The closed generator has been heated and the results are presented in Table 3. Comparing
with the results of Section 6.2.1, it could be deduced that hi improve by conduction through the
gas when keq increases with concentration.

6.2.3. Open generator


We have also compared the experimental and the computed results in the case of an open
generator. This experiment has been done in the temperature range ½20; 70C, and under a
pressure P ¼ 10:2 bars. We find: keq ¼ 0:43 W/m K and hi ¼ 33:45 W/m2 K. The experimental
results confirm well good the validity of the proposed model, as shown in Fig. 3.

7. CPC–HP–generator assembly model

The analysis of the obtained result concerning the porous medium thermal conductivity keq and
the heat exchange coefficient, hi between the wall and the porous medium shows that the thermal
transfer is very poor. To overcome this technical problem, a system of a generator/HP assembly is
studied. The HP itself is coupled with a CPC as shown in Fig. 4. The most important HP design
consideration in this case is the amount of power the HP is capable of transferring at a given
operational temperature Tv and HP size (i.e.; length and external radius) as CPC parameters should
also be taken into account by the mean of concentration ratio, Cc. However, if the HP is driven
beyond its capacity, its effective thermal conductivity will be significantly reduced. Table 4 gives
some HP’s and CPC’s parameters. The viscous, entrainment, sonic and boiling limits are much
greater in comparison with the capillary limit and so does not suppose any limit problem [17–19].
Moreover, the great advantage of combining solar collector with HP is the reduction of the
collector area [20] and greater power production [21], other interesting consequences could arise
from combining the HP with an adsorptive generator.

7.1. CPC–HP model

For a single glass CPC–HP, the following assumptions are made:

• the glass, reflector, and heat pipe surface have uniform temperature distribution at a given
time;
2548 F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555

Fig. 4. The solar adsorption machine.

• the CPC component properties are independent of temperature;


• the heat flow through the CPC is one dimensional.

We express only the energy balance equation relative to the HP area (energy equations relative
to the reflector and the glass could be deduced)
oThp 
mhp chp ¼ Shp Iu þ Sre hr;ðre;hpÞ ðTre  Thp Þ þ Sgl hr;ðgl;hpÞ Tgl  Thp þ Shp hcv;ðgl;hpÞ ðTgl  Thp Þ
ot
þ Shphcv ;ðre;hpÞ ðTre  Thp Þ ð7Þ

7.1.1. Convective, conductive and radiative heat transfer coefficient


General correlation for the variation of internal convective heat transfer in low-concentration-
ratio CPCs, considering the effect of inclination, latitudinal and tracking configuration of the CPC
are given in [22,23]. The radiative, conductive and global heat transfer coefficients were defined
and modelled in a manner outlined in [24].
F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555 2549

Table 4
Some CPCs and HPs parameters
Parameter Symbol Units Value
Glass transmittance sgl 0.9
HP’s container absorptance ahp 0.95
Reflector reflectance qore 0.9
Average number of reflections, at the reflector, ni 1.38
for irradiation which reaches the HP
Concentration ratio Cc 1.45
Reflector height H m 12.5 · 102
Reflector width W m 22 · 102
HP’s vapour radius rv m 0.017
Wick porosity ewick % 73.3
Wick nature ss 200
Wick thickness ewick m 2 · 103
HP’s adiabatic length Lad m 0.1
HP’s evaporator length Lev m 1
HP’s condenser length Lcon m 0.25
HP’s transported power Q kW 0.680

The overall and optical efficiencies adopted to the CPC under consideration could be given in
the form
Thp  Tamb
g ¼ gopt  htot ð8Þ
Igb
where
Iu
gopt ¼  sgl  ahp  qoni
re
Igb
and
ID
Iu ¼ IB þ
Cc
IB and ID are beam and diffuse radiation, respectively.

7.2. HP–generator

HP-reactor model is similar to that of Section 5. Assumptions made are identical to those
specified in Section 2. However, reactor pressure is considered non-uniform. During the heating
and cooling phases with closed reactor, the total mass of ammonia remains constant and is the
sum of the total adsorbed mass and the free mass of gaseous ammonia. Therefore, the rate of
change in ammonia mass over time can be expressed as:
R
d ðma ðT ðrÞ; P Þ þ mg ðT ðrÞ; P ÞÞ
dr ¼ 0 ð9Þ
dt
where mg is the gaseous ammonia mass (kg/kg carbon).
2550 F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555

The boundary conditions are given below.

7.2.1. Boundary conditions

7.2.2. Heating phase


oT
r ¼ r1 ; keq ¼ hgb;m ðThp  T Þ
or
oT
r ¼ r2 ; ¼ 0 ðthermal insulationÞ
or
hgb;m is the mean overall heat transfer coefficient between the HP’s vapour space and the gener-
ator, and is given by
hgb;c þ hgb;h
hgb;m ¼ ð10Þ
2
where
r
1 1 Ln rhp;e
hp;i 1 1
¼ þ þ ¼ ð11Þ
hgb;h Shp;i hhp;i Shp;i 2pkss l hge;i Shp;e hgb;c Shp;e

7.2.3. Cooling phase

oT
r ¼ r1 ; ¼0
or
oT
r ¼ r2 ; ¼0
or

8. Results and discussion

For HP–generator model solving, we apply the finite difference method (implicit scheme). The
equivalent thermal conductivity, keq and generator’s internal heat transfer coefficient hg;i , are those
identified experimentally. The distribution of temperature, pressure and adsorbed mass are cal-
culated for each step of time. The global heat transfer coefficient between the HP’s wall and the
generator, hgb;m is 34.73 W m2 k1 . An east–west aligned and symmetric CPC is considered. The
determination of temperature profile on the CPC–HP’s surface is done by an iterative process.
The simulation is done for Lleida city, in Spain, on the summer season, considering a random
august day. Fig. 5 shows the horizontal global radiation variation and the temperature profile on
the heat pipe’s evaporator surface. The surface collector is only 0.22 m2 and the random daily
received solar radiation is 7 kW m2 day1 of which only 5.6 kW m2 day1 corresponds to the
heating phase. The optimal inclination, hopt , is found to be equal to 0, i.e. horizontal. CPCs and
HPs parameters are given in Table 4. The temperature profile on the heat pipe’s evaporator
surface reaches maximal value of 137 C. The HPs radial heat flux and the wick’s effective thermal
F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555 2551

Fig. 5. Horizontal global solar radiation and temperature profile on the heat pipe’ s evaporator surface variation.

conductivity are 5.6 kW m2 and 1.12 W m1 K1 , respectively. The temperature drop through
both the HP’s container and wick is therefore of 10 C. It could be deduced that the heat pipe
operates from 10.42 to 15.2 h (i.e.; 4 h 47 min) giving a mean power of 680 W at a mean tem-
perature of 100 C to the generator. The generator is heated (resp. cooled) from Tads to Tge
(respectively, Tge to Tads ). When the pressure in the generator reaches the condenser pressure
(respectively, evaportar’s pressure), valve 1 (respectively, valve 2) opens and the adsorbate and the
condenser (respectively, evaporator) are communicated. Condensation–desorption phase
(respectively, evaporation–adsorption phase) starts. Following characteristics temperatures:
Tamb ¼ ½19; 29:8 C, Tev ¼ 0 C, Tcon ¼ 30 C have been adopted for calculations. The generator’s
internal radius, external radius and length are 0.019, 0.06 and 0.25 m respectively. The generator
contains 5 kg of activated carbon. The chosen HPs working fluid is water because of its good
latent heat of vaporisation and also for its compatibility with stainless steel 304 itself compatible
with ammonia. Fig. 6 shows that generator’s temperature reaches its maximal temperature of 100
C at 15.2 h and that the desorption-condensation phase is of about of 3 h. The generator pressure
increases from the initial value of 4.2 to 10.1 bar in only 1.75 h (i.e.; from 10.42 to 12.17 h), Fig. 7,
when the adsorbed mass is kept constant at the value of 0.24 kg kg1 . This pressure remains
constant until the end of heating, leading to a strong decreasing in the adsorbed mass to reach 0.1
kg kg1 . During the cooling phase, the pressure decreases to reach the initial value of 4.2 bar in
10 h. At this stage the adsorbed mass remains constant at the value of 0.1 kg kg1 . No pressure
variation is observed and pressure stabilise at the value of 4.2 bar when the adsorbed mass

Fig. 6. Generator temperature variation during the day.


2552 F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555

Fig. 7. Generator pressure and adsorbed mass variation during the day.

increase to reach the initial value of 0.24 kg kg1 corresponding to a generator temperature of 23
C. Exergy destruction or exergy loss, EDe ¼ Td sgen , where Td is the dead state temperature taken
equal to Tamb and sgen is the entropy generation owing to internal irreversibilities, is present over
the whole refrigeration process except when phase change occurs. The entropy generation, sgen , is
computed by the following equation:
 
QC gHg QC þ gHg
sgen ¼  þ 
Tev Tge Tamb
found by combining the first and the second law of thermodynamics. QC is the cold produc-
tion and Hg is the global daily radiation on horizontal plane. The coefficient of performance,
COP, defined as the ratio of cooling to energy input is strongly related to the adimensional exergy
loss
EDe
AEL ¼
Qc
the adimensional generator’s temperature
Tge
Tge ¼
Tamb
and adimensional evaporator’s temperature
Tev
Tev ¼
Tamb
as
1
1
T
COP ¼   ge
1
 1 þ AEL
Tev
The solar coefficient of performance COPs can be deduced therefore as
COPs ¼ g  COP
F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555 2553

Fig. 8. AEL, COP and COPs variation as a function of the ambient temperature.

Fig. 8 shows that COPs variation for different ambient temperatures. The minimal value corre-
sponding to Tamb ¼ 19 C is 0.1426 and the maximal value corresponding to Tamb ¼ 24:2 C is
0.144. COPs is then nearly constant with changing ambient temperature. This implies that g is
inversely proportional to COP. In fact, COP is an increasing function of Tge which is inversely
proportional to g. On the other hand, and as was expected, COP (respectively, AEL) is a
decreasing (respectively, increasing) function of ambient temperature and it could be deduced
therefore, that the nominal point of the machine corresponds to Tamb ¼ 24:9 C, AEL ¼
COPs ¼ 40% and COPs ¼ 14.37%.

9. Conclusion

From both experimental and numerical results corresponding to heat and mass transfer into the
double casing generator, both equivalent thermal conductivity keq and internal heat transfer
coefficient hi are well identified. These values are used after in the HP–generator assembly model.
Following conclusions can be drawn from the simulation of the whole solar installation:

• The proposed design allows for reaching the generator’s initial temperature and therefore the
initial adsorbed mass. This implies that similar performance will be achieved for the following
cycles, which is not the case for a direct generator-solar collector assembly.
• The design allows also for using many generators of different size as a dismantled jacket. The
internal radius r1 is fix while r2 and Lge could be variables ðLcon 6 Lge 6 Lcon þ Lad Þ. The quantity
of the produced cold could be therefore adjusted. In this case, the transferred power to the gen-
erator will increase from one design to another as the transferred power (i.e; wicking limitation)
is inversely proportional to an effective HP length, Leff equal to the distance from the midpoint
of the evaporator to the midpoint of the condenser.
• The great amount of power of the HP within a small vapour space (Qmean ¼ 680 W, SV ¼ 9
cm2 ), produces an increase of HP’s wall temperature and improves clearly the heat transfer
from the primary solar energy to the generator.
• The COP, AEL and COPs study, shows the competitivity of the proposed design. As a result,
COPs ¼ 14.37%, Tamb ¼ 24:9 C, Tev ¼ 0 C, Tcon ¼ 30 C and Tge ¼ 100 C.
2554 F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555

The CPC–HP–generator assembly constitute an interesting flexible design. Thermal perfor-


mance of the machine could be improved by varying design parameters and/or using finned HP’s
condenser. Both studies will be done at a later stage.

References

[1] L. Luo, M. Feidet, R. Boussehain, Etude thermodynamique de machine a adsorption, Entropie 183 (1994) 3–
11.
[2] R.E. Critoph, Performance limitation of adsorption cycles for solar cooling, Int. J. Solar Energy 41 (1986) 21–
31.
[3] R.E. Critoph, Activated carbon adsorption cycles for refrigeration and heat pumping, Int. J. Carbon 27 (1989) 63–
70.
[4] A. Al mers, Ph. D. Thesis, Etude du transfert de chaleur et de masse dans un lit fixe de charbon actif reagissant par
adsorption avec de l’ammoniac, Faculte des Sciences de Tetouan, July 2002.
[5] L.M. Sun, F. Meunier, B. Mishler, Etude analytique des distributions de temperatures et de concentration a
l’interieur d’un grain spherique d’adsorbant solide soumis a un echelon de pression de vapeur adsorbable, Int. J.
Heat Mass Tran. 29 (1986) 1393–1406.
[6] A. Kanolt, Etude experimentale et theorique de l’adsorption non-isotherme de vapeur d’eau sur des grains isoles de
zeolithe 4a , Euromech 194 (1985) 113–114.
[7] D.M. Ruthven, L.-K. Lee, Kinetics of non-isothermal sorption: systems with bed diffusion control, Aich. J. 27
(1981) 654–663.
[8] K. Chihara, M. Suzuki, K. Kawazoe, Effect of heat generation on measurement of adsorption rate by gravimetric
method, Chem. Eng. Sci. 31 (1976) 505–507.
[9] J.J. Guilleminot, F. Meunier, J. Pakleza, Heat and mass transfer in a non-isothermal fixed bed solid adsorbent
reactor, Int. J. Heat Mass Trans. 30 (1987) 1595–1606.
[10] L.W. Wang, J.Y. Wu, R.Z. Wang, Y.X. Xu, S.G. Wang, X.R. Li, Study of performance of activated carbon–
methanol adsorption systems concerning heat and mass transfer, Appl. Therm. Eng. 23 (2003) 1605–1617.
[11] R.E. Critoph, L. Turner, Heat transfer in a granular activated carbon beds in the presence of adsorbable gases, Int.
J. Heat Mass Trans. 38 (1995) 1577–1585.
[12] R.E. Critoph, S.J. Metcalf, Specific cooling power intensification limits in ammonia–carbon refrigeration systems,
Appl. Therm. Eng. 24 (2004) 661–678.
[13] Chemviron, Cranular activated carbon, Chemviron, Bruxelles, 1988.
[14] Institut International de froid., Tables et diagrammes pour l’industrie du froid, Proprietes thermodynamiques du
R12, R22, R717, Paris, 1981.
[15] A. Mahamane, Etude de l’adsorption de vapeurs pures sur solides poreux, These de doctorat, Universite de Mons,
1988.
[16] B.P. Bering, M.M. Dubinin, V.V. Serpinsky, Theory of volume filling for vapour adsorption, J. Col. Int. Sci. 21
(1966) 378–393.
[17] P.D. Dunn, D.A. Reay, Heat pipes, Pergamon Press, Oxford, 1994.
[18] G.P. Peterson, An Introduction to Heat Pipes Modelling, Testing, and Applications, John Wiley and Sons, New
York, 1994.
[19] A. Faghri, Heat Pipes Sciences and Technology, Taylor and Francis, London, 1995.
[20] F. Aghbalou, A. Touzani, M. Mada, A cylindrical parabolic solar collector heat pipe assembly, in: Proceedings of
The International Thermal Energy And Environment Congress, Marrakech, Morocco, vol. 2, 1997, pp. 867–872.
[21] F. Aghbalou, A. Touzani, M. Mada, M. Charia, A. Bernatchou, A parabolic solar collector heat pipe heat
exchanger reactor assembly for cyclohexane’s dehydrogenation: a simulation study, Int. J. Renewable Energy 14
(1–4) (1998) 60–67.
[22] P.C. Eames, B. Norton, Detailed parametric analysis of heat transfer in CPC solar energy collectors, Solar Energy
50 (4) (1993) 321–338.
F. Aghbalou et al. / Applied Thermal Engineering 24 (2004) 2537–2555 2555

[23] A.F. Kothdiwala, B. Norton, P.C. Eames, The effect of variation of angle of inclination on their performance of
low-concentration-ratio compound parabolic concentrating solar collectors, Solar Energy 55 (4) (1995) 301–309.
[24] D.E. Prapas, B. Norton, S.D. Probert, Thermal design of compound parabolic concentrating solar energy
collectors, J. Solar energy Engng 109 (1987) 161–168.

You might also like