Sunteți pe pagina 1din 34

Journal of Enhanced Heat Transfer, 23(4):315–348 (2016)

NUMERICAL INVESTIGATION OF THE


EFFECTS OF THE NUMBER OF RADIAL
LONGITUDINAL FINS ON THE MELTING OF
PARAFFIN WAX IN A CYLINDRICAL
ANNULUS
Ashish Agarwal

Department of Mechanical Engineering, Technocrats Institute of


Technology-Excellence, Bhopal, MP, India, E-mail:
er ashishagarwal@yahoo.com

A numerical simulation of the melting process of a phase-change material (PCM) in a horizontal


cylindrical annulus has been studied with and without heat transfer enhancement techniques. In this
study, longitudinal fins were used to enhance heat transfer in a cylindrical annulus. A numerical
study was carried out for melting of paraffin wax, with and without fins, using the Fluent finite-
volume code. Three configurations of a cylindrical annulus with a constant wall temperature of the
inner pipe were studied: (1) without fins, (2) with four fins, and (3) with eight fins. To analyze the
thermal behavior of the system, the results of the presented numerical simulation are used to show the
movement of the melting front in the cylindrical annulus for the three aforementioned configurations.
The results show that comparatively high heat transfer was achieved in the cases with eight and four
fins compared to the plain annulus: 90% and 73% melt fractions were easily achieved in 1000 and
900 s in the cases of a cylindrical annulus with eight and four fins, respectively, and the heat transfer
and melting rate become slower after theses time limits. To generalize the results, two dependent
dimensionless parameters are used: The Nusselt number (Nu) and melt fraction of the PCM. The
product of the Stefan and Fourier numbers (FoSte) takes into account the transient phase change and
heat conduction and serves as an independent dimensionless parameter. The trends presents in the
dimensional analysis are a step toward generalization and can be used in the design of PCM-based
heat storage systems.

KEY WORDS: extended surface, phase-change material (PCM), passive enhance-


ment, thermal energy storage

1. INTRODUCTION

In recent years, most of the developing countries around the world are facing an en-
ergy problem because of the large gap between the demand and supply of energy. This
problem can be eliminated to some extent by utilizing renewable energy sources such as
solar energy. Solar energy is available abundantly throughout the world; however, it is
not continuous and its intensity also varies over time. Due to the aforementioned reason,

1065-5131/16/$35.00
c 2016 by Begell House, Inc. www.begellhouse.com 315
316 Agarwal

NOMENCLATURE

Amush mushy zone constant Tref reference temperature


[kg/(m3 /s)] (◦ C or K)
A(γ) porosity function Ts solid temperature (◦ C or K)
cp specific heat (J/kg · K) t time (s)
Fo Fourier number (= αt/l2 ) ui velocity component in the
h specific enthalpy (J/kg) i direction (m/s)
href reference enthalpy (J/kg) uj velocity component in the
hs sensible enthalpy (J/kg) j direction (m/s)
k thermal conductivity xi Cartesian coordinate
(W/m · K) xj Cartesian coordinate
L latent heat of material (J/kg)
l characteristic length (m) Greek Symbols
Nu Nusselt number α thermal diffusivity of the
(= Ql/∆T k) phase-change material
P pressure (Pa) (m2 /s)
Q mean heat flux (W/m2 ) β thermal expansion
Si source term in the coefficient (K−1 )
momentum equation γ liquid fraction
Ste Stefan number (= cp ∆T /L) ∆ difference
Tf melting temperature µ dynamic viscosity
(◦ C or K) (N · s/m2 )
Tl liquid temperature ρf density of the phase-change
(◦ C or K) material (kg/m3 )

the acceptability and reliability of solar-based thermal systems is lower than conven-
tional systems. A properly designed heat storage system increases the reliability of a
solar thermal system by bridging the gap between the energy demand and availability.
Phase-change materials (PCMs) are used to store heat in the form of latent heat. Most
PCMs do not fulfill the criteria for an ideal storage medium. Most of them have low
thermal conductivity and the problem of supercooling (Zalba et al., 2003; Agarwal and
Sarviya, 2016).
Many researchers have reported the thermal and heat transfer characteristics of la-
tent heat storage systems with different geometrical configurations during charging and
discharging. Khodadadi and Zhang (2001) numerically studied the melting process of
a PCM in a spherical container. Their results showed that the rate of melting is faster
at the top region of a sphere than at the bottom region. They investigated the effect of
convection on the melting rate.

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 317

Hosseini et al. (2012) conducted an experimental and computational study of the


melting behavior of a medium temperature PCM in a horizontal shell and tube heat
exchanger. Their results showed that the melting front appeared at different times at po-
sitions closer to the heat transfer fluid (HTF) tube, progressing at different rates outward
in the shell. It was also observed that the total melting time was decreased to 37% by
increasing the inlet water temperature to 80◦ C.
Seeniraj et al. (2002) investigated numerically the transient behavior of high-tempe-
rature PCMs during melting in a shell and tube heat exchanger. They observed the effect
of thin circumferential fins on the melting rate of the PCM inside the horizontal shell.
Complete melting of the PCM was not observed in the case of an unfinned tube. Some
quantity of the PCM at the exit of the tube remained in the solid state. They also reported
that complete melting was observed in the axial direction by the addition of annular fins.
Due to the annular fins, there was a high temperature difference between the HTF and
the PCM’s melting point.
Agarwal and Sarviya (2016) experimentally evaluated the performance of shell and
tube heat storage during the melting and solidification process using air as the HTF. The
results showed that the discharging time was greater compared to the charging time, and
the charging time increased by 20% when the temperature of the HTF decreased from
90◦ C to 80◦ C.
Since most PCMs have low thermal conductivity, this leads to less effective heat
transfer during charging and discharging of heat storage media. Hence, heat transfer en-
hancement techniques are required in order to bring paraffin-based PCMs to successful
applications. This problem is addressed through the insertion of metal fins, matrix, foam,
graphite matrix, etc. The use of metal fins with different geometries has been proposed in
the literature to improve the heat transfer. Numerous experimental and numerical studies
have been conducted to investigate the consequences of using fins with different geome-
tries in the melting and solidification process (Rathod and Banerjee, 2015; Bentilla et
al., 1966; Hoover et al., 1971; Humphries, 1974; Griggs et al., 1974; Abhat, 1976; De
Jong and Hoogendoorn, 1981).
Rozenfeld et al. (2015) studied melting of a PCM in a horizontal double pipe concen-
tric storage unit with three longitudinal fins attached to an inner tube. A laboratory-scale
transparent unit was constructed to observe the melting of the PCM inside the pipe.
Close-contact melting was achieved by supplying heat to the outer shell of the unit. A
numerical model was developed that included gravity-induced rotational motion of the
solid, primary melting on a vertical fins, and secondary melting at the shell. The numeri-
cal modeling was validated by the experimental results, and then used for parametric and
dimensional analyses. The results of the study showed that the melt fraction depends on
the Fourier and Stefan numbers.
Ogoh and Groulx (2012) conducted a numerical study to evaluate the effect of the
number of longitudinal fins on the storage characteristics of a vertical cylindrical latent
heat energy storage system (LHESS) during the charging process. The numerical study

Volume 23, Issue 4, 2016


318 Agarwal

was performed for different numbers of fins ranging from 1 to 27. The results showed
that after 12 h the amount of energy stored increases nearly linearly with the addition
of fins up to 12 fins; further addition of fins increased the total energy stored by smaller
amounts. Hosseinizadeh et al. (2011) conducted experimental and numerical investiga-
tions to evaluate the effectiveness of fins in a PCM-based heat sink. The operating vari-
ables considered during the study were the fin height, number of fins, fin thickness, and
power level. An appreciable increase in the overall thermal performance was observed
by increasing the number of fins and the fin height. The performance of the heat sink was
greatly improved up to an optimal fin thickness; no further improvement was observed
beyond the optimum fin thickness.
Hosseini et al. (2015) examined the effect of longitudinal fins on the heat transfer in
a double pipe heat exchanger containing a PCM during the charging process. The effect
of the fin height and Stefan number on the thermal performance of the heat exchanger
was studied. The results of the study showed that a longitudinal fin leads to less melting
time and deeper heat penetration. The heat absorbed during the initial charging period
is a function of the fin’s height; increasing the Stefan number leads to a higher melting
rate.
Al-Abidi et al. (2013) numerically evaluated the performance of a triplex tube
heat storage unit consisting of three concentric tubes. The HTF flow was in the in-
ner and outer tubes, and the PCM was stored in the middle tube. The longitudinal fins
were attached to the inner and outer tubes to increase the heat transfer rate. The
melting time and melt fraction of the heat storage unit were compared for
two, four, and eight longitudinal fins. The computational results showed that the shor-
test melting time was observed for the heat storage unit with eight fins. For the
same heat storage configuration, Al-Abidi et al. (2014) also performed an experimen-
tal investigation. The effect of the HTF mass flow rate and inlet temperature on the
charging process was examined. The results of the study claimed that the HTF inlet tem-
perature had a greater influence on the PCM melting process than did the HTF mass flow
rate.
Liu and Groulx (2014) experimentally investigated the heat transfer inside a cylin-
drical latent heat storage unite with longitudinal fins during the charging and discharging
process. The longitudinal fins were attached to a central copper pipe to enhance the heat
transfer rate. Dodecanoic acid was used as the PCM in this study. Two configurations
of longitudinal fins (i.e., straight fins and angle fins) were studied and compared. It was
found that for both configurations conduction was dominated by the heat transfer mode
during the beginning of the charging process. The total melting time for the angled fins
was lower compared with the straight fins for low HTF inlet temperatures. However, for
a relatively high HTF inlet temperature there was no difference in the heat transfer effect
between the straight and longitudinal fins.
Murray and Groulx (2014a) experimentally evaluated the performance of a vertical
LHESS with water as the HTF during charging and discharging. The proposed system

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 319

was designed to be used with a solar domestic hot water system. The longitudinal fins
were attached to the HTF pipe to enhance the heat transfer. The heat transfer and phase-
change behavior of the PCM inside the vertical LHESS during consecutive charging
and discharging process was studied. The results showed that the longitudinal fins were
effective in increasing the heat transfer by improving the natural convective flow of the
melt inside the cylinder. The effect of natural convection was found to be significant dur-
ing melting. Murray and Groulx (2014b) also evaluated the performance of the system
during simultaneous charging and discharging and found that the direct heat transfer be-
tween the heat source (hot HTF) and heat sink (cold HTF) only occurred when the PCM
melted. The low thermal conductivity of the PCM limits the heat transfer when the PCM
is a solid material.
Sciacovelli et al. (2014) proposed tree-shaped fins to enhance the performance of a
shell-and-tube latent heat thermal energy storage unit. The shape of the fins was opti-
mized in the transient condition to improve the heat transfer and melting time through
the combined use of computational fluid dynamics modeling and the response surface
method. Y-Shaped fins with wide angles between branches are preferable for short op-
erating times whereas smaller angles are necessary for long operating times. The results
showed that the system efficiency increased by 24% due to optimization of the fin shape
in the storage unit.
Sharifi et al. (2011) developed a numerical model to simulate the melting of a PCM
housed within an internally finned metal enclosure. They reported on the influence of
the number of fins, fin length and thickness, and hot wall temperature on the melting
process.
Mat et al. (2013) numerically investigated the melting process in a triplex-tube heat
exchanger with PCM RT82. They developed a two-dimensional (2D) numerical model
using the Fluent 6.3.26 (Ansys, Inc., USA) software program. In their paper, internal,
external, and internal–external fin enhancement techniques were studied to improve the
heat transfer between the PCM and HTF.
Li and Liu (2013) conducted an experimental study to evaluate the effects of metal
thermal conductivity enhancers on the thermal performance of PCMs. Two PCMs (paraf-
fin wax RT25 and RT42) and four metal thermal conductivity enhancers (vertical fin,
volume fraction = 1.8%; horizontal fin, volume fraction = 1.8%; honeycomb structure,
volume fraction = 2.7%; and square cell structure, volume fraction = 3.6%) were stud-
ied. The experimental results showed that the metal thermal conductivity enhancers im-
proved the heat transfer greatly even with a small volume fraction (< 4%), especially for
a PCM with lower melting temperature.
Manglik and Jog (2016) proposed a dry-cooling (air-cooling) system in place of a
conventional evaporative cooling system for large-scale thermoelectric power plants to
save on the consumption of water and energy. The system consists of a novel daytime
peak-load shifting system that precools the day-time peak-temperature air in order to
mitigate the concomitant second-law limitation for air cooling. The highly compact and

Volume 23, Issue 4, 2016


320 Agarwal

enhanced air pre-cooler transfers the day-time heat load to a unique thermal energy stor-
age unit. The thermal energy storage unit operates over a range of temperatures and is
recharged by an asynchronous night-time air-cooled exchanger. The proposed system
increases the output of the power plant and makes dry air cooling viable.
Heat transfer in a cylindrical thermal energy storage system represents a transient
phase-change problem. Heat transfer in such a problem is nonlinear due to the moving
solid–liquid interface. In most of previous research studies, the phase-change problem
is solved by enthalpy–porosity methods and temperature-based equivalent heat capacity
methods. The enthalpy–porosity method is adopted in this work to model the combined
convection–diffusion phase change of the PCM. This method has been used in numerical
solutions by several researchers, such as Gartling, (1980), Voller et al. (1987), Hunter
and Kuttler (1989), Brent et al. (1988), Cao and Faghri (1990), and Zivkovic and Fujii
(2001).
The literature review shows that the selection of fins to enhance the heat transfer
depends on the orientation, shape, and size of the heat storage container. The effective-
ness of the fin was studied by varying the configuration (triplex, concentric, etc.) and
the orientation of the heat storage container (vertical, horizontal, etc.). Apart from these
parameters, the shape of the fins (i.e., circular, longitudinal, etc.) also affects the heat
transfer enhancement. Different fin shapes (i.e., circular, longitudinal, and plate fins)
have been used in the literature to enhance heat transfer, and their effectiveness has been
evaluated by numerical and experimental studies. However, in the previous studies, the
effect of the number of longitudinal fins on heat transfer enhancement in a horizontal
heat storage container has not been studied in detail. Some studies evaluated the effect
of fins on heat transfer enhancement in heat sinks (Sharifi et al., 2011), horizontal cylin-
drical heat storage units (Hosseini et al., 2012, 2015; Seeniraj et al., 2002; Agarwal and
Sarviya, 2016; Rozenfeld et al., 2015; Liu and Groulx, 2014), vertical cylindrical heat
storage units (Ogoh and Groulx, 2012; Murray and Groulx, 2014a,b), and triplex tube
heat storage units (Al-Abidi et al., 2013, 2014; Mat et al., 2013) using thin circumfer-
ential fins (Seeniraj et al., 2002) and longitudinal fins (Rozenfeld et al., 2015; Ogoh and
Groulx, 2012; Hosseini et al., 2015; Al-Abidi et al., 2013, 2014; Liu and Groulx, 2014;
Murray and Groulx, 2014a,b).
The present numerical study aims to compare the heat transfer during melting of a
PCM in a cylindrical annulus based on three geometrical cases, i.e., without fins and
with four and eight longitudinal fins. In this context, the primary objective of the present
study is to investigate the effect of the fins on the natural convection and heat transfer
during melting of the PCM in the annulus of a heat exchanger. The effect of varying the
number of fins on the melting and heat transfer behavior in a comprehensive way is the
prime difference between this paper and similar previous studies. In the present research
work, the transient thermal behavior of a latent heat storage system with and without
fins during melting is analyzed and presented at different instants of time in term of melt
contours.

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 321

2. COMPUTATIONAL DOMAIN AND BOUNDARY CONDITIONS

The computational domains under examination are shown in Fig. 1. The heat storage
unit is considered as a 2D model. The experimental study conducted by Hosseini et

(a)

(b)

FIG. 1.

Volume 23, Issue 4, 2016


322 Agarwal

(c)

FIG. 1: (a–c) Schematic illustrations of the configurations of the computational domains


of the plain annulus, the finned cylindrical annulus with four fins, and the finned cylin-
drical annulus with eight fins

al. (2012) on a horizontal shell and tube latent heat storage unit shows that variation
of the paraffin wax temperature in the axial direction is very small such that the 2D
approximation can be easily adopted in the present work. This model was selected to
reduce the computational time. The geometry used in this study is composed of an inner
copper pipe having a diameter of 22 mm. The diameter of the outer pipe is 120 mm.
The PCM is filled between the outer and inner pipe. Three cases are considered in the
present study: a plain annulus (case 1), a finned cylindrical annulus with four fins (case
2), and a finned cylindrical annulus with eight fins (case 3). In all cases, the fins are
equally spaced and the radial length of the fins is 43 mm. Copper was selected as the
fin material and the fin thicknesses selected were 2 and 1 mm for the cylindrical annulus
with four fins and the cylindrical annulus eight fins, respectively. The thickness of the
fins was selected such that the total mass of the PCM in the annulus remains constant in
the cylindrical annulus with four and eight longitudinal fins. The inner copper pipe works
as a HTF pipe. A constant wall temperature is applied at the wall of the inner pipe. The
PCM is initially at a temperature of 303 K. The outer wall is considered adiabatic and
the temperature of the inner pipe is maintained at a constant temperature of 363 K. The
computational domain is discretized using quadrilateral mesh elements. The resulting
system of nonlinear partial differential equations is solved.
A paraffin wax produced by the Indian Oil Corporation (New Delhi, India) was con-
sidered as the PCM in the present study. Differential scanning calorimetry (DSC) anal-
ysis of the paraffin wax sample was conducted to determine its specific heat, latent heat,

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 323

and melting range. The DSC analysis was performed on a Thermal Analyzer Pyris DSC
6000 (TA Instruments, PerkinElmer, Inc., USA) in a heating and cooling cycle. The re-
sult of the DSC analysis is presented in Fig. 2. The thermo-physical properties of the
PCM are listed in Table 1.

3. NUMERICAL PROCEDURE

The semi-implicit method for pressure-linked equations (SIMPLE) algorithm (Patankar,


1980) was utilized to solve the governing equations. The quadratic upstream interpola-
tion for convective kinematics (QUICK) differencing scheme (Ferziger and Peric, 2002)
was used to solve the momentum and energy equations, whereas the pressure staggering
option (PRESTO) scheme (Ferziger and Peric, 2002) was adopted to solve the pressure
correction equation. By solving the governing equations at each time step, the liquid
mass fraction was updated using the enthalpy–porosity equation.
Different grid elements, such as square and triangular shapes were tried in this study.
Since the geometry is not very complicated, a square-element grid structure was selected

FIG. 2: Result of the DSC analysis

TABLE 1: Thermo-physical properties of paraffin wax


Property Value
Melting temperature range (K) 314–328
Latent heat capacity (kJ/kg) 176
Specific heat (kJ/kg · k) 2.8
Density (kg/m3 ) 835
Thermal conductivity (W/m · k) 0.21

Volume 23, Issue 4, 2016


324 Agarwal

in the present study. The results were obtained from different grid densities, such as 50,
80, and 100 k. These results were compared and the 80 k grid distribution was chosen
in this study. The time step in the calculations was as small as 0.1 s and the number of
iterations for each time step was 25. The number of iterations for every time step fixed at
25 was found to be sufficient to satisfy the convergence criteria. Time step independency
tests were carried out to select the optimum time step in the present study. The optimum
time step generates accurate results and saves computational time in the transient study.
Three time steps were selected in the present study, namely, 0.05, 0.1, and 0.5 s. A time
step of 0.1 s was selected to save computational time and obtain accurate results. The
grid size and the time step were chosen after careful examination of the independency of
the results. The simulations were considered to converge when the scaled residuals were
lower than 10−3 , 10−3 , and 10−6 for continuity, velocity, and energy, respectively. The
under-relaxation factors for the pressure correction, density, body forces, momentum,
and liquid fraction were equal to 0.3, 1, 1, 0.7, and 0.9, respectively. The numerical
approach makes it possible to calculate the processes that occur inside the solid PCM
(conduction) and liquid PCM (convection), and to take into account the phase change,
moving boundary due to variation in the PCM volume, and solid phase motion in the
melt.

4. MATHEMATICAL FORMULATION

In order to simulate the phase change in a PCM in a cylindrical annulus, the enthalpy–
porosity method (Brent et al., 1988; Gong et al., 1999) is used. In this technique, the
liquid melt fraction in each cell is computed every iteration, based on enthalpy balance.
The mushy zone is the region where the porosity increases from 0 to 1 as the PCM melts.
When the region is completely solid, the porosity is zero and the flow velocity in that
zone also drops to zero.
The enthalpy of the PCM is defined as the sum of the sensible enthalpy (hs ) and the
latent heat (∆H)
h = hs + ∆H (1)

The sensible enthalpy is computed as


Z T
hs = href + cp dT (2)
Tref

where href is the reference enthalpy at the reference temperature Tref , and cp is the spe-
cific heat.
The change in enthalpy due to the phase change is given by

∆H = γL (3)

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 325

where γ = liquid fraction, and L = latent heat of the material. If the phase change occurs
over a range of temperatures Ts < T < Tl , then according to the enthalpy content of the
material the liquid fraction can be calculated by the following relations:
∆H
γ= =0 if Ts < T (4)
L
∆H T − Ts
γ= = if Ts < T < Tl (5)
L Tl − Ts
∆H
γ= = 1 if T > Tl (6)
L
The flow in the molten PCM is considered to be laminar, incompressible, and 2D. The
viscous dissipation term is considered negligible, such that the viscous incompressible
flow and the temperature distribution in the annulus space are described by the Navier–
Stokes and thermal energy equations, respectively. The change in volume during the
phase-change process is neglected. The governing equations used in this study are given
as follows:

• Continuity


(ρui ) = 0 (7)
∂xi

• Momentum

∂ ∂ ∂ 2 ui ∂P
(ρui ) + (ρui uj ) = µ − + ρgi + Si (8)
∂t ∂xi ∂xi xj ∂xi

• Energy
 
∂ ∂ ∂ ∂T
(ρh) + (ρui h) = K (9)
∂t ∂xi ∂xi ∂xi

where h is the specific enthalpy of the PCM; ρ is the density; T is the temperature; k
is the thermal conductivity, Si is the source term; ui and uj are the velocity compo-
nents in the i and j directions, respectively; and xi and xj are the Cartesian coordinates,
respectively.
The source term in the momentum equation is given by the following equation:

Amush 1 − γ2

Si = −A (γ) ui = ui (10)
γ3 + ε

Volume 23, Issue 4, 2016


326 Agarwal

where Amush is the mushy zone constant, which reflects the morphology of the melt-
ing front. The value of Amush ranges between 104 and 108 kg/m3 · s. The value of
Amush = 108 kg/m3 · s was taken in this study, which had been suggested by Shmueli
et al. (2010). Here, ε is a small number to avoid division by zero and its value is taken
as being equal to 0.001 in this study, and A(γ) is defined as the porosity function, which
governs the source term in the momentum equation and is defined based on the Carman–
Kozeny relation for a porous medium introduced by Brent et al. (1988). The value of this
function depends on the melt fraction. This function goes to zero since γ is equal to 1
and becomes large when γ approaches zero. This function changes the velocity in the
computational cells during the phase change.
The change in the density of the PCM (in the liquid phase) with temperature is
calculated by the Boussinesq approximation (Shatikan et al., 2005) and is given by
ρf
ρ= (11)
β (T − Tf ) + 1

where Tf is the melting temperature; ρf is the density of the PCM at the melting temper-
ature; and β is the thermal expansion coefficient. The value of β = 0.001 K−1 is chosen
based on the analysis of the detailed data presented by Humphries and Griggs (1977).
The change in density within the liquid phase that derives from natural convection is
considered only in the body force terms in the momentum equation. Following Reid et
al. (1987), the dynamic viscosity of the liquid PCM is expressed as
 
1790
µ = 0.001 × exp −4.25 + (12)
T

5. VALIDATION OF THE NUMERICAL RESULTS


5.1 Validation of the Numerical Model for the Phase-Change Process
In order to evaluate the ability of the present numerical model to model the phase-change
process, a test case was considered. The test case consists of melting of pure gallium in a
rectangular cavity, as performed by Gau and Viskanta (1986). This test case was solved
numerically by the present numerical model using the same computational domain, ini-
tial conditions, and thermophysical properties as used in Gau and Viskanta (1986). The
temperature of the left wall of the rectangular cavity was kept at a constant level in heat-
ing the PCM. The top and bottom walls were assumed to be thermally insulated. The
results are shown in the form of movement of the melting front with time in 2D compu-
tational domains. The distance in the Y direction in Fig. 3 shows the distance from the
bottom of the rectangular cavity in the Y direction. The distance in the X direction in
Fig. 3 shows the distance from the left heated wall in the X direction. The movement of
the melting front is shown in 2D coordinates (X–Y ) after 3, 6, and 10 min of melting

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 327

FIG. 3: Comparison of movement of the melting front from the present study with the
results reported by Brent et al. (1988) and Gau and Viskanta (1986)

of pure gallium in a rectangular cavity. The results of the present model were compared
with the experimental and numerical results obtained by Gau and Viskanta (1986) and
Brent et al. (1988), respectively. As shown in Fig. 3, acceptable agreement was found be-
tween the results obtained from the present numerical code and those obtained in Brent
et al. (1988) and Gau and Viskanta (1986).

5.2 Experimental Setup and Procedure


Double pipe latent heat storage was fabricated to validate the numerical model of the
phase-change process of the PCM in the cylindrical annulus. Figure 4 shows a schematic
diagram of the experimental setup, which includes the double pipe heat storage, resis-
tance heating element, proportional/integral/derivative (PID) controller, data logger, and
dual core PC. Double pipe latent heat storage consists of concentric tubes of 500 mm
length. The diameters of the outer and inner tubes were 120 and 22 mm, respectively.
The inner tube is the heat transfer tube and the outer tube is insulated with glass wool
(8 mm thick) to reduce heat loss to the surrounding area.

Volume 23, Issue 4, 2016


328 Agarwal

FIG. 4: Schematic diagram of the experimental setup

A data logger is used to measure and record the temperature at definite intervals of
time from different locations of the annulus. Nine K-type thermocouples were inserted
in the PCM at 15 mm intervals in the radial and angular directions, as shown in Fig. 5.
These thermocouples were used to measure the temperature in the radial and angular
directions of the annulus of the heat storage.

FIG. 5: Locations of thermocouples in the cylindrical annulus

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 329

The resistance heating element was inserted in the inner tube and the paraffin wax
was filled into the annulus of the heat storage. The resistance heating element was used
to supply heat to the air, which was used as the HTF in the present study. To maintain the
temperature of the air at a constant level at the inlet of the HTF tube, the PID temperature
controller was used in this study. Heat is transferred to the HTF tube by the air, as it
travels through the HTF tube and its temperature changes along the length of tube. A
temperature gradient also exists along the length of HTF tube. The PID controller is
a control loop feedback mechanism (controller) commonly used in industrial control
systems to maintain the temperature at a constant level.
The temperature controller is an instrument generally used to control the tempera-
ture. The temperature controller compares the process temperature with a set temperature
value (desired value), and the difference between these values is known as the deviation
or error. The temperature controller controls the power input to the heating element based
on these errors to bring the process temperature to the desired value. The output signals
from the temperature controller are finally fed to the final control element to remove or
inject the heat.
A Selec PID controller (DTC324) is coupled to the resistance heating element to
maintain the temperature of the air (HTF) at a constant level. The Selec DTC324 (Selec
Controls Pvt. Ltd., Navi Mumbai, India) is an auto-tuning PID controller. The output of
this controller is in the form of the PID (or ON/OFF). A solid-state relay is connected to
the output of the PID controller to receive output signals from it and to control the power
supply to heat the resistive heat element. The response of the PID controller to control
the temperature to the set value is shown in Fig. 6. The set point for the PID controller is
363 K and the resistance heating element was initially at atmospheric temperature. The
heat is supplied to air until its measured temperature exceeds the set point. The error
signals in the beginning were high, which were reduced over time. The three controls
of the PID controller (i.e., proportional, integral, and derivative) reduce the error signals
over time and the air temperature becomes stable after 12 min, as shown in Fig. 6.

5.3 Validation of the Numerical Model by the Experimental Results

To validate the results of the numerical simulation for melting the PCM in the cylindrical
annulus the results of the present code were compared with the experimental results.
During the experimental run, the temperature gradient existed along the length of the
HTF tube due to temperature variations of the HTF and axial conduction of heat in the
HTF tube. However, in the experimental setup the temperature of the PCM was measured
at a 50 mm distance from the HTF tube inlet, where the temperature of the HTF tube was
nearly equal to the HTF inlet temperature of 90◦ C. For validation, a numerical simulation
was conducted at a HTF pipe temperature of 90◦ C. The temperature at different angular
locations (0◦ , 90◦ , and 180◦ ) of the PCM near the HTF tube inlet was measured by the
thermocouples inserted in the annulus of the heat storage near the inlet. The average

Volume 23, Issue 4, 2016


330 Agarwal

FIG. 6: The response of the PID controller for keeping the air at constant temperature

temperatures of the PCM at 0◦ , 90◦ , and 180◦ angular locations of the annulus obtained
by the numerical and experimental work are shown in Fig. 7. It is clearly shown in
Fig. 7 that the temperature at the 90◦ and 180◦ angular locations is always higher than
the temperature at the 0◦ angular location due to the formation of convective current.
Figure 7 shows that the results of the numerical simulation are in good agreement with
the experimental data.

6. RESULTS AND DISCUSSION

A computer simulation study was performed in the present work for a plain annulus, a
cylindrical annulus with four fins, and a cylindrical annulus with eight fins (cases 1, 2,
and 3) subjected to an inner cylinder wall temperature of 90◦ C. The instantaneous con-
tours of the melt fraction were obtained by computational analysis at different instants
of time during melting of the PCM in cylindrical annulus. The aim of present study was
to analyze the effect of fins on the melting process of a PCM in a cylindrical annulus.

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 331

FIG. 7: Comparison of timewise variations of the average temperature in different an-


gular directions (0◦ , 90◦ , and 180◦ ) obtained by numerical and experimental studies of
the plain cylindrical annulus

6.1 Melting of the PCM in a Plain Cylindrical Annulus (Case 1)

The melt fraction contours show the position of the melt front at different instants of time
during melting. Figure 8 shows the melt fraction contours for an inner wall temperature
of 90◦ C for a plain cylindrical annulus. The temperature of inner wall is maintained at
90◦ C, which is 36◦ C higher than the melting temperature of the PCM. The melt contour
is concentric to the inner cylinder wall in the beginning due to conduction mode of
heat transfer. During the initial period of melting the solid starts to melt initially due to
conduction and the melting front is symmetrical in all directions. On further melting,
natural convection comes into play and the shape of the melting front becomes irregular,
which is clearly visible in Fig. 8 at the time instant of 500 s. The melt contour at time
1000 s shows that the melting front becomes wider in the uppermost portion compared
with lower portion. As the melting time progresses natural convection intensifies, a wider
melting region forms at the upper region of the annulus, and molten PCM occupies the
major part of the annulus. Melting in the uppermost section is complete at the time
instant of 2500 s; after this time period the melting rate becomes very slow. It can be

Volume 23, Issue 4, 2016


332 Agarwal

Fig. (a) Melt Contour at 100 sec Fig. (b) Melt Contour at 500 sec Fig. (c) Melt Contour at 1000 sec

Fig. (d) Melt Contour at 1500 sec Fig. (e) Melt Contour at 2000 sec Fig. (f) Melt Contour at 2500 sec

Fig. (g) Melt Contour at 3000 sec Fig. (h) Melt Contour at 3500 sec Fig. (i) Melt Contour at 4000 sec

FIG. 8: (a–i) Melt fraction contours in the PCM at different instants of time for a plain
cylindrical annulus
seen that as time advances from 2500 to 27,000 s, the melting front remains stagnant in
the lower section. The melting front moves in a downward direction at a very slow rate.
The melting rate in the lower section is significantly lower than any other part of the
annulus.
The transient variation of the melt fraction for the plain annulus is shown in Fig. 9.
A rapid rise in the melt fraction is observed at the beginning of melting due to the
buoyancy-driven melting of the PCM in the upper section of the annulus. A higher melt-
ing rate is observed up to the time instant of 3500 s, as shown in Fig. 9. The melt fraction

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 333

FIG. 9: Transient variations of the melt fraction for the plain annulus

at the time instant of 3500 s is 65%. Complete melting was achieved in the upper portion
at this time instant and 35% of the PCM remained in a solid state in the bottom region.
As time advances the melting rate becomes slower due to the conduction-dominated
mode of heat transfer in the bottom region.

6.2 Melting of the PCM in a Finned Cylindrical Annulus with Four Fins
(Case 2)

A computational study on the concentric plain annulus shows that there is a low melting
rate in the bottom portion of the annulus. As shown in Fig. 8, the melting occurs faster in
the upper section of the annulus compared to lower section due to the high temperature
in the upper section. The melting front is created near the inner pipe wall and moves in
an upward direction due to the effect of the buoyancy forces. The mushy region, which is
bound by the liquid (55◦ C) and solid (41◦ C) isotherms, is concentric to the cylinder wall
in the beginning and becomes wider in the uppermost section as the melting proceeds.
From the numerical results for a plain cylindrical annulus, it was found that the melting
front becomes horizontal in the lower region at the time instant of 4000 s and moves in

Volume 23, Issue 4, 2016


334 Agarwal

a downward direction at a very slow rate. The melting in the bottom of the annulus is
dominated by conduction. Due to the conduction-dominated zone in the bottom portion,
the temperature of the lower section increases very slowly. Due to aforementioned reason
the plain cylindrical annulus is a less efficient heat storage device.
To solve this problem, four longitudinal fins were attached to the inner cylindrical
pipe wall to improve the heat transfer rate. These fins were placed at 90◦ angular spacing
on the inner pipe wall. The thicknesses of fins were 2 mm and 3 mm gaps left between
the tip of the fins and the outer cylinder. The geometry of the fins was selected such
that the total volume of the PCM in the annulus remained constant for the three different
cases considered in the present study. The cross-sectional view of the cylindrical annulus
considered in this case is shown in Fig. 1(b). The melting process of the PCM in the
cylindrical annulus is observed by the melt fraction contours for different instants of
time during the melting process. The melt fraction contours show the position of the
melt front in the annulus. To simulate the melting in the finned cylindrical annulus, all
of the parameters in the present study were similar to those used in the plain cylindrical
annulus study.
Figure 10 shows the melt contours in the finned cylindrical annulus with four fins at
different instants of time. The four fins were attached to the outer surface of the inner
cylinder to increase the heat transfer area. It was observed that the heat transfer rate was
very low in the case of the plain cylindrical annulus due to the low thermal conductivity
of the PCM.
The aim in considering this case was to analyze the effect of the fins on the melting
rate and natural convective flow of the molten PCM. Initially, the PCM melted near
the surface of the inner cylinder and fins due to conduction heat transfer. As the time
advanced, more melting was observed near the fins and the wall of the inner pipe, and
convection heat transfer came into effect. At the time instant of 200 s more melting was
observed around the inner pipe and at the end of fin located at θ ∼180◦ . This was a
result of the high rate of heat transfer due to mixed conduction–convection heat transfer.
Buoyant forces developed along the surface of the fin located at θ ∼180◦ , and as a result
less dense molten PCM moved from the bottom to the top end of the fins and became
trapped at the top end of the fin; thus, more melting was observed at this location.
As time advanced, the melting area became wider at the end of fin compared to
other area near the fin, as shown in Fig. 10(a) at the time instant of 200 s. Melting was
uniform along the fin located at θ ∼90◦ . More melting was observed in the top region
compared to the bottom region. It is evident from Fig. 10(c) that the melting rate is high
in the upper region compared to lower region at 600 s, where the melting front takes
the shape of a parabola due to the blockage effect produced between the fins at θ ∼90◦
and ∼180◦ in the top region of the annulus. Intense buoyant forces at the top portion
led to a higher melting rate. At this time instant two recirculation regions were found in
the upper region; one smaller recirculation region was located at the end of the fin and
a larger recirculation region was located near the pipe wall. A distorted melting front

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 335

Fig. (a) Melt Contour at 200 sec Fig. (b) Melt Contour at 400 sec Fig. (c) Melt Contour at 600 sec

Fig. (d) Melt Contour at 800 sec Fig. (e) Melt Contour at 1000 sec Fig. (f) Melt Contour at 1200 sec

Fig. (g) Melt Contour at 1400 sec Fig. (h) Melt Contour at 1600 sec Fig. (i) Melt Contour at 1800 sec

Fig. (j) Melt Contour at 1800 sec Fig. (k) Melt Contour at 2000 sec Fig.(l) Melt Contour at 2200 sec

FIG. 10: (a–l) Melt fraction contours in the PCM at different instants of time for a finned
cylindrical annulus (four fins)

Volume 23, Issue 4, 2016


336 Agarwal

was observed at time 800 s in the upper region and a smaller solid region was left in
the upper region. By comparing the melt contours at 600 and 800 s, it was found that
the melting front moved at a very slow rate in the lower region of the annulus and the
shape of melting front remained parabolic with time. It is evident from the melt contours
at the time instant of 1000–2200 s that the velocity of the melting front is progressively
reduced with time and the melting front remains stagnant at the bottom region of the
annulus. The transient variation of the melt fraction for the finned cylindrical annulus
with four fins is shown in Fig. 11, where the trend of the melt fraction profile is similar
to the plain annulus but melting is completed comparatively in much less time compared
to the plain annulus.

6.3 Melting of the PCM in a Finned Cylindrical Annulus with Eight Fins
(Case 3)

In this section, the numerical results of melting of the PCM in a cylindrical annulus with
eight fins are presented. Eight fins were attached to the wall of an inner cylinder at a fixed
angular position. The fins were placed at 45◦ angular spacing on the inner pipe wall. A

FIG. 11: Transient variations of the melt fraction for a finned cylindrical annulus (four
fins)

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 337

3-mm gap was left between the tip of the fins and the outer cylinder. A schematic diagram
of the cylindrical annulus with eight fins is shown in Fig. 1(c). The presented study
focused on the development of a convection-dominated zone in the bottom section of
the annulus. To simulate melting in the finned cylindrical annulus, all of the parameters
in the present study were similar to the plain cylindrical annulus study. Figure 12 shows
the melt fraction contours of the finned cylindrical annulus with eight fins at different
instants of time.
Figure 12(a) shows that the melting fronts are parallel to the fin surface in the initial
stage of melting (100 s) and more melting is observed in the lower region near the inner

Fig. (a) Melt Contour at 100 sec Fig. (b) Melt Contour at 200 sec Fig. (c) Melt Contour at 300 sec

Fig. (d) Melt Contour at 400 sec Fig. (e) Melt Contour at 500 sec Fig. (f) Melt Contour at 600 sec

Fig. (g) Melt Contour at 700 sec Fig. (h) Melt Contour at 800 sec Fig. (i) Melt Contour at 900 sec

FIG. 12.

Volume 23, Issue 4, 2016


338 Agarwal

Fig. (j) Melt Contour at 1000 sec Fig. (k) Melt Contour at 1100 sec Fig. (l) Melt Contour at 1200 sec

Fig. (m) Melt Contour at 1300 sec Fig. (n) Melt Contour at 1400 sec Fig. (o) Melt Contour at 1500 sec

FIG. 12: (a–o) Melt fraction contours in PCM at different instants of time for a finned
cylindrical annulus (eight fins)

pipe wall due to the buoyancy effect, which develops due to the temperature gradient
in the molten PCM. As more melting takes place the melting front is no longer parallel
to fin surface, as shown in Fig. 12(b). The melting front takes the shape of a parabola
between the two consecutive fins in the lower region at the time instant of 200 s.
High convective flow of molten PCM was observed in the bottom region near the
pipe wall due to buoyant forces. More melting was observed near the pipe wall in the
bottom region at 300 s; however, this was not observed in the top region. Melting took
place in this region along the radial direction parallel to the fin surface, as shown in the
Fig. 12(d) at 400 s; after 700 s less convective flow is observed, which leads to a slower
melting rate. The melting front is still in the shape of parabola in the lower region. The
parabolic shape is produced due to the blockage effect developed by the fins. Complete
melting is achieved in the upper region at 600 s; 78% melting is achieved at this time
instant, and 12% of the PCM is in the solid phase in the bottom portion. It is evident
from Fig. 12 that the melting front propagates at a very slow rate in the bottom region.
Approximately 940 s is required for melting 12% of the PCM existing in the bottom
portion. The transient variation of the melt fraction for the finned cylindrical annulus
with eight fins is shown in Fig. 13. Melting of the PCM is complete after 2040 s, which
is much less time compared to the plain annulus and the finned cylindrical annulus with
four fins.

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 339

FIG. 13: Transient variations of the melt fraction for an inner wall temperature of 90◦ C
for a finned cylindrical annulus (eight fins)

7. COMPARISON OF TRANSIENT VARIATIONS OF THE MELT FRACTION


IN A PLAIN ANNULUS, A FINNED CYLINDRICAL ANNULUS WITH FOUR
FINS, AND A FINNED CYLINDRICAL ANNULUS WITH EIGHT FINS

The transient variations of the melt fraction in a plain annulus, a finned cylindrical annu-
lus with four fins, and a finned cylindrical annulus with eight fins are shown in Fig. 14.
Considerable improvement in the heat transfer rate is observed by attaching fins to the
wall of the inner pipe. Complete melting of the PCM is observed in the cylindrical annu-
lus with four fins and the cylindrical annulus with eight fins at time durations of 3365 and
2040 s, respectively. However, the total melt fraction in the case of the plain cylindrical
annulus is only 68% at the time instant of 4000 s.
In all three cylindrical annulus cases considered in this study it was observed that
the melting rate becomes stagnant after a given time duration. This time instant is called
the threshold time. The threshold times for the plain cylindrical annulus, the cylindrical
annulus with eight fins, and the cylindrical annulus with four fins were 2500, 1100, and
900 s, respectively. Comparatively, a low heat transfer rate was observed in the case of

Volume 23, Issue 4, 2016


340 Agarwal

FIG. 14: Transient variations of the melt fraction for the plain annulus, the finned cylin-
drical annulus with four fins, and the finned cylindrical annulus with eight fins

the plain cylindrical annulus, where 57% of the melt fraction was achieved at the time
instant of 2500 s. High convective flow was observed in the upper portion of the annulus,
such that complete melting was achieved easily in the upper portion; a very long period
of time was required to melt the PCM in the bottom portion of the annulus.
In the case of cylindrical annulus with four fins, 74% of the melt fraction was ob-
served at the time instant of 900 s. High heat transfer due to high convective flow of
the melt was observed up to this time instant. At this time instant complete melting was
observed in the upper portion and 26% of the PCM in the bottom region remained in a
solid state.
In the case of cylindrical annulus with eight fins, 93% of the melt fraction was ob-
served at the time instant of 1100 s. After this time duration the melting rate became
slower, as shown in Fig. 14. Some portion (12%) of the PCM remained in a solid state
in the bottom region of the annulus. Complete melting of this portion of the PCM was
difficult to achieve due to natural convective flow and melting that transferred most of
the heat to the upper region of the annulus. Approximately 940 s was required to melt
12% of the PCM, which existed in the bottom portion.

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 341

8. DIMENSIONAL ANALYSIS
The results of the study discussed previously shows that the melting process depends
on the number of fins in the system. To generalize the results, dimensional analysis was
applied to the present study. The dimensionless Fourier Number (Fo) is used to deal with
the transient heat conduction problem. The Fourier Number is defined as Fo = αt/l2 ,
where α is the thermal diffusivity of the PCM, t is the time, and l is the characteristic
length. The value of the Fourier number varies from 0.02 to 5.7, 0.009 to 1.15, and
0.03 to 2.3 for the plain annulus, the annulus with four fins, and the annulus with eight
fins, respectively. In the phase-change problem, the Fourier number is not sufficient to
generalize the results because it cannot take into account the phase-change process. For
this reason, the Stefan number (Ste) is also involved in the analysis. In our study the
Stefan number is defined as Ste = cp ∆T /L, where cp is the specific sensible heat of the
PCM, ∆T is the difference between the temperature of inner pipe wall and the mean
melting temperature of the PCM, and L is the latent heat of melting. In the present
study, the Stefan number was 0.67 for the three cases, i.e., the plain annulus, the finned
cylindrical annulus with four fins, and the finned cylindrical annulus eight fins. The
product of the Stefan and Fourier numbers (FoSte) takes into account the transient phase
change and heat conduction and serves as an independent dimensionless parameter.
To generalize the results, the two dependent dimensionless parameters are used: the
Nusselt number (Nu) and the melt fraction of the PCM. The Nusselt number is defined
as Nu = Ql/∆T k, where Q is the mean heat flux at the interface, ∆T is the temperature
difference between the inner heat transfer wall and the PCM mean melting temperature,
k is the thermal conductivity of the PCM, and l is the characteristic length.
The results of the study are shown in Figs. 15 and 16 in a dimensionless form. Figures
15 and 16 present the Nusselt numbers and melt fractions, respectively, for cases 1,
2, and 3, i.e., the plain annulus, the finned cylindrical annulus with four fins, and the
finned cylindrical annulus with eight fins, respectively. The results show that the Nusselt
number and melt fraction depend on the product of the Fourier and Stefan numbers.
Figure 15 shows the variations of the Nusselt number as a function of the product
of the Fourier and Stefan numbers for cases 1, 2, and 3 considered in our study. The
value of the Nusselt number varies from 3 to 0, 9 to 0, and 8.8 to 0, for cases 1, 2,
and 3, respectively. The Nusselt number had a maximum value for cases 2 and 3 in the
beginning, where the temperature difference is the maximum value between the heat
transfer surface and the PCM. In case 1, the Nusselt number increased first and then
decreased. In case 1, as the amount of molten PCM increased with time in the upper half
of the cylinder, the natural convection effects were strengthened, which further increased
the heat transfer. Since the melting of the PCM was complete in upper half of cylinder,
the convective mode of the heat transfer became weaker with time and the conductive
mode of heat transfer dominated. Such behavior was not observed in cases 2 and 3, where
fins were present and mixed conductive and convective heat transfer were responsible for
melting of the PCM.

Volume 23, Issue 4, 2016


342 Agarwal

FIG. 15: The Nusselt number versus the product of the Fourier and Stefan numbers for
cases 1, 2, and 3

The application of the Fourier and Stefan number discussed previously led to the
generalized results presented in Fig. 16, where the melt fraction is shown versus FoSte
for cases 1, 2, and 3. The number of fins has a strong influence on the melting. In all
cases the melt fraction increased rapidly initially up to some critical value of FoSte, and
then the rate of change in melt fraction became low. One can see from Fig. 16 that the
trends of the curves for the melt fraction are similar for the plain annulus (case 1), the
finned cylindrical annulus with four fins, and the finned cylindrical annulus with eight
fins (cases 2 and 3) when the value of the melt fraction is relatively high (up to about
0.8).
At lower melt fractions, the trend is slightly different for the plain annulus (case 1).
The curves diverge slightly at lower melt fractions for cases 2 and 3 compared to case 1,
and the curves continue to follow nearly the same pattern for cases 2 and 3. As one can
see, in all cases the melting rate becomes weaker at a higher value of the melt fraction.
In case 1 the melt fraction rate becomes very slow after 0.6 because after this period the

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 343

FIG. 16: Melt fractions for cases 1, 2, and 3 versus the product of the Fourier and Stefan
numbers

convective mode of heat transfer becomes very weak and further transfer of heat to the
PCM is purely governed by the conductive mode of heat transfer. Some of the findings in
the present study will be restricted to the specific geometry (i.e., the horizontal cylinder
with longitudinal fins) due to gravity and volumetric expansion of the PCM. The trends
presented in the dimensional analysis are a step toward generalization. This work will
continue in a future study.

9. FIN EFFECTIVENESS

The heat transfer enhancement due to the fin is calculated by the term fin effectiveness.
The fin effectiveness is the ratio of heat transfer with fins and without fins. This term
indicates the extra amount of heat being transferred by the fin. Fin effectiveness is rep-
resented by the following equation:

Volume 23, Issue 4, 2016


344 Agarwal

Qfinned
εfin = (13)
Qunfinned
Figure 17 shows the variations of fin effectiveness for the four and eight fins with time.
The fin effectiveness decreases with time. In the case of a plain annulus the heat transfer
in the initial stage of melting is very low due to conductive-dominated heat transfer. The
majority of heat is transferred by conduction. In the case of a finned cylindrical annulus
the heat transfer in the initial stage of melting is high compared to a plain annulus due
to enhancement of thermal conductivity by longitudinal fins. Thus, the fin effectiveness
is high in the initial stage of melting. As more and more melting of the PCM occurs,
natural convective flow of the molten PCM forms in the melted region, which leads to
convective-dominated heat transfer. The fin effectiveness decreases with time due to an
increase in convective heat transfer. The effectiveness of the fin is high in the initial stage
of melting due to conductive-dominated heat transfer.

10. CONCLUSIONS
In the present work, a computational study was performed to investigate the thermal and
heat transfer characteristics during melting of latent heat storage. Numerical investiga-
tions were carried out on a cylindrical annulus subjected to constant wall temperature
using the finite-volume code Fluent. The numerical investigations were carried out for
conductive and buoyancy-driven convective heat transfer in a cylindrical annulus during

FIG. 17: Variations of fin effectiveness for four and eight fins

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 345

the melting process, with and without heat transfer enhancement techniques. Three types
of PCM-filled annuli were studied namely, a plain cylindrical annulus, a cylindrical an-
nulus with four longitudinal fins, and a cylindrical annulus with eight longitudinal fins.
The developed computational model was able to track the movement of the melting front
at different instants of time during melting. The results of the numerical study revealed
the effect of the fins on the melting rate. Based on the findings of the computational
study of the melting of paraffin wax in a cylindrical annulus, the following conclusions
have been drawn:

1. Buoyancy-driven convective melting was observed in the uppermost section of the


annulus and melting in the lower portion was dominated by conduction in the case
of the plain annulus.

2. It was found that melting in the plain annulus was dominated by heat conduction
followed by free convection; melting of paraffin wax starts from the center of the
annulus, moves in a radial direction in the upper part of the annulus, and finally
moves from the top to the bottom portion.

3. The results show that a sharp rise in temperatures takes place in the uppermost
section of the heat storage because of the buoyancy effects.

4. Insertion of longitudinal fins is advantageous to increase the heat transfer in all


portions of the annulus. A considerable increase in the melting rate was observed
in the case of the finned cylindrical annulus compared to the plain cylindrical
annulus.

5. Approximately 70% of the PCM in the annulus melted in comparatively much less
time compared to the total melting time in the case of finned cylindrical annulus
(four and eight fins) at a wall temperature of 90◦ C. For the remaining 30% of
the PCM in the lower portion of the annulus, melting took much less time for
the finned cylindrical annulus with eight fins compared to the finned cylindrical
annulus with four fins at the same temperature of inner wall. Comparatively, high
heat transfer due to mixed conduction and convection leads to quick melting of
the major part of the PCM in the lower portion of the annulus.

6. Comparatively, a slow heat transfer rate was observed in the case of the plain
cylindrical annulus. Approximately 57% of the melt fraction was easily achieved
in a moderate amount of time. High convective flow was observed in the upper
portion of the annulus, such that complete melting was achieved easily in the upper
portion. A very long time period was required to melt the PCM in the bottom
portion of the annulus.

Volume 23, Issue 4, 2016


346 Agarwal

7. The longitudinal finned system with eight fins is recommended for double pipe
heat storage with paraffin wax as the PCM because it achieved higher heat transfer
and melting rate.

8. The results of the present study broaden the insight into the phenomena of the
PCM melting process in a horizontal cylinder with longitudinal fins. The trends
presented in the dimensional analysis are a step toward generalization and can be
used in the design of PCM-based heat storage systems.

REFERENCES
Abhat, A., Experimental investigation and analysis of a honeycomb-packed phase change mate-
rial device, Proc. of AIAA 11th Thermophysics Conference, Paper AIAA-76-437, 1976.
Agarwal, A. and Sarviya, R.M., An experimental investigation of shell and tube latent heat stor-
age for solar dryer using paraffin wax as heat storage material, Eng. Sci. Technol. Int. J., vol.
19, no. 1, pp. 619–631, 2016.
Al-Abidi, A.A., Mat, S., Sopian, K., Sulaiman, M.Y., and Mohammad, A., Internal and external
fin heat transfer enhancement technique for latent heat thermal energy storage in triplex tube
heat exchangers, Appl. Therm. Eng., vol. 53, pp. 147–156, 2013.
Al-Abidi, A.A., Mat, S., Sopian, K., Sulaiman, M.Y., and Mohammad, A., Experimental study
of melting and solidification of PCM in a triplex tube heat exchanger with fins, Energy Build.,
vol. 68, pp. 33–41, 2014.
Bentilla, E.W., Sterrett, K.F., and Karre, L.E., Research and development study on thermal con-
trol by use of fusible materials, Interim Report NSL-65-16-1, Northrop Space Laboratories,
NASA Marshall Space Flight Center, Huntsville, AL, 1966.
Brent, A.D., Voller, V.R., and Reid, K.J., Enthalpy-porosity technique for modeling convection-
diffusion phase change: Application to the melting of a pure metal, Numer. Heat Transfer, vol.
13, no. 3, pp. 297–318, 1988.
Cao, Y. and Faghri, A., A numerical analysis of phase change problem including natural convec-
tion, J. Heat Transfer, vol. 112, pp. 812–815, 1990.
De Jong, A.G. and Hoogendoorn, C.J., Improvement of heat transport in paraffins for latent heat
storage systems, Proc. of International TNO Symposium on Thermal Storage of Solar Energy,
Amsterdam, The Netherlands: Martinus Nijhoff Publishers, pp. 123–133, 1981.
Ferziger, J.H. and Peric, M., Computational Methods for Fluid Dynamics, Berlin: Springer, 2002.
Gartling, D.K., Finite element analysis of convective heat transfer problems with change of
phase, Computer Methods in Fluids, K. Morgan, C. Taylor, and C.A. Brebbia, Eds., London:
Pentech, pp. 257–284, 1980.
Gau, C. and Viskanta, R., Melting and solidification of a pure metal on a vertical wall, J. Heat
Transfer, vol. 108, no. 1, pp. 174–181, 1986.
Gong, Z.X., Devahastin, S., and Mujumdar, A.S., Enhanced heat transfer in free convection-
dominated melting in a rectangular cavity with an isothermal vertical wall, Appl. Therm. Eng.,
vol. 19, no. 12, pp. 1237–1251, 1999.

Journal of Enhanced Heat Transfer


Effect of Radial Longitudinal Fins on Melting of Paraffin Wax 347

Griggs, E.I., Pitts, D.R., and Humphries, W.R., Transient analysis of a thermal storage unit in-
volving a phase change material, Proc. of ASME Winter Annual Meeting, Paper 74-WA/HT-21,
1974.
Hoover, M.J., Grodzka, P.G., and O’Neill, M.J., Space thermal control development, Final Report
LMSCHREC D225500, Lockheed Huntsville Research and Engineering Center, Huntsville,
AL, 1971.
Hosseini, M.J., Ranjbar, A.A., Rahimi, M., and Bahrampoury, R., Experimental and numerical
evaluation of longitudinally finned latent heat thermal storage systems, Energy Build., vol. 99,
pp. 263–272, 2015.
Hosseini, M.J., Ranjbar, A.A., Sedighi, K., and Rahimi, M., A combined experimental and com-
putational study on the melting behavior of a medium temperature phase change storage ma-
terial inside shell and tube heat exchanger, Int. Commun. Heat Mass Transfer, vol. 39, no. 9,
pp. 1416–1424, 2012.
Hosseinizadeh, S.F., Tan, F.L., and Moosania, S.M., Experimental and numerical studies on per-
formance of PCM-based heat sink with different configurations of internal fins, Appl. Therm.
Eng., vol. 31, pp. 3827–3838, 2011.
Humphries, W.R., Performance of finned thermal capacitors, NASA Technical Note D-7690,
NASA, Washington, DC, 1974.
Humphries, W.R. and Griggs, E.I., A design handbook for phase change thermal control and
energy storage devices, NASA Technical Paper 1074, NASA Scientific and Technical Infor-
mation Office, Hampton, VA, 1977.
Hunter, L.W. and Kuttler, J.R., The enthalpy method for heat conduction problems with moving
boundaries, J. Heat Transfer, vol. 111, no. 2, pp. 239–242, 1989.
Khodadadi, J.M. and Zhang, Y., Effects of buoyancy-driven convection on melting within spher-
ical containers, Int. J. Heat Mass Transfer, vol. 44, pp. 1605–1618, 2001.
Li, Y. and Liu, S., Effects of different thermal conductivity enhancers on the thermal perfor-
mance of two organic phase-change materials: Paraffin wax RT42 and RT25, J. Enhanced
Heat Transfer, vol. 20, no. 6, pp. 463–473, 2013.
Liu, C. and Groulx, D., Experimental study of the phase change heat transfer inside a horizon-
tal cylindrical latent heat energy storage system, Int. J. Thermal Sci., vol. 82, pp. 100–110,
2014.
Manglik, R.M. and Jog, M.A., Resolving the energy-water nexus in large thermoelectric power
plants: A case for application of enhanced heat transfer and high-performance thermal energy
storage, J. Enhanced Heat Transfer, vol. 23, no. 4, pp. 263–282, 2016.
Mat, S., Al-Abidi, A.A., Sopiana, K., Sulaimana, M.Y., and Mohammada, A.T., Enhance heat
transfer for PCM melting in triplex tube with internal–external fins, Energy Convers. Manage.,
vol. 74, pp. 223–236, 2013.
Murray, R.E. and Groulx, D., Experimental study of the phase change and energy characteris-
tics inside a cylindrical latent heat energy storage system: Part 1 consecutive charging and
discharging, Renewable Energy, vol. 62, pp. 571–581, 2014a.

Volume 23, Issue 4, 2016


348 Agarwal

Murray, R.E. and Groulx, D., Experimental study of the phase change and energy characteris-
tics inside a cylindrical latent heat energy storage system: Part 2 simultaneous charging and
discharging, Renewable Energy, vol. 63, pp. 724–734, 2014b.
Ogoh, W. and Groulx, D., Effects of the number and distribution of fins on the storage char-
acteristics of a cylindrical latent heat energy storage system: A numerical study, Heat Mass
Transfer, vol. 48, no. 10, pp. 1825–1835, 2012.
Patankar, S.V., Numerical Methods in Heat Transfer and Fluid Flow, Washington, DC: Hemi-
sphere Publishing Corporation, 1980.
Rathod, M.K. and Banerjee, J., Thermal performance enhancement of shell and tube latent heat
storage unit using longitudinal fins, Appl. Therm. Eng., vol. 75, pp. 1084–1092, 2015.
Reid, R., Prausnitz, J., and Poling, B., The Properties of Gases and Liquids, New York: McGraw-
Hill, 1987.
Rozenfeld, T., Kozak, Y., Hayat, R., and Ziskind, G., Close-contact melting in a horizontal cylin-
drical enclosure with longitudinal plate fins: Demonstration, modeling and application to ther-
mal storage, Int. J. Heat Mass Transfer, vol. 86, pp. 465–477, 2015.
Sciacovelli, A., Gagliardi, F., and Verda, V., Maximization of performance of a PCM latent heat
storage system with innovative fins, Appl. Energy, vol. 137, pp. 707–715, 2014.
Seeniraj, R.V., Velraj, R., and Narasimhan, N.L., Thermal analysis of a finned-tube LHTS module
for a solar dynamic power system, Heat Mass Transfer, vol. 38, nos. 4-5, pp. 409–417, 2002.
Sharifi, N., Bergman, T.L., and Faghri, A., Enhancement of PCM melting in enclosures with
horizontally-finned internal surfaces, Int. J. Heat Mass Transfer, vol. 54, pp. 4182–4192, 2011.
Shatikan, V., Ziskind, G., and Letan, R., Numerical investigation of a PCM-based heat sink with
internal fins, Int. J. Heat Mass Transfer, vol. 48, pp. 3689–3706, 2005.
Shmueli, H., Ziskind, G., and Letan, R., Melting in a vertical cylindrical tube: Numerical investi-
gation and comparison with experiments, Int. J. Heat Mass Transfer, vol. 53, pp. 4082–4091,
2010.
Voller, V.R., Cross, M., and Markatos, N.C., An enthalpy method for convection/diffusion phase
change, Int. J. Numer. Methods Eng., vol. 24, no. 1, pp. 271–284, 1987.
Zalba, B., Marı́n, J.M., Cabeza, L.F., and Mehling, H., Review on thermal energy storage with
phase change: Materials, heat transfer analysis and applications, Appl. Therm. Eng., vol. 23,
no. 3, pp. 251–283, 2003.
Zivkovic, B. and Fujii, I., An analysis of isothermal phase change of phase change material
within rectangular and cylindrical containers, Sol. Energy, vol. 70, pp. 51–61, 2001.

Journal of Enhanced Heat Transfer

S-ar putea să vă placă și