You are on page 1of 52

PII: S1359-4311(16)34418-0

DOI: http://dx.doi.org/10.1016/j.applthermaleng.2016.12.127
Reference: ATE 9747

To appear in: Applied Thermal Engineering

Received Date: 11 August 2016


Revised Date: 20 December 2016
Accepted Date: 27 December 2016

Please cite this article as: Y. Kozak, M. Farid, G. Ziskind, Experimental and comprehensive theoretical study of
cold storage packages containing PCM, Applied Thermal Engineering (2016), doi: http://dx.doi.org/10.1016/
j.applthermaleng.2016.12.127

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Experimental and comprehensive theoretical study

of cold storage packages containing PCM

Yoram Kozak1*, Mohammed Farid2 and Gennady Ziskind1

1
Department of Mechanical Engineering, Ben-Gurion University of the Negev, Beer-Sheva,
Israel, Phone: 972-8-6477089, Fax: 972-8-6472813, e-mail: kozaky@post.bgu.ac.il
2
Department of Chemical and Materials Engineering, The University of Auckland, Auckland,
New Zealand, Phone: 649-3737599 Fax: 649-3737463, e-mail: m.farid@auckland.ac.nz

*Corresponding author

1
ABSTRACT

The current research explores both experimentally and theoretically the thermal

performance of transported insulated cold storage packages that can keep products at a low

temperature. In order to sustain the product cold as long as possible, it is suggested to

combine a conventional insulation with a phase-change material (PCM), thus taking

advantage of its latent heat. A number of useful modelling approaches are suggested and

tested for an insulation-PCM system. First, a rather simplified, but still preserving the main

physical traits of the problem, analytical model is developed. This model reveals the

dimensionless groups that govern the problem. It also shows the existence of an optimal

insulation thickness that maximizes the melting time of the PCM, and thus keeps the product

at a low temperature for a longer period. Then, a more exact, yet fast and computationally

inexpensive, numerical model is introduced. This conduction-based model, which takes into

account the three-dimensional structure of the insulation and natural convection, is validated

against experiments performed on two packages, with different dimensions and PCMs. Both

experimental and numerical results show that the heating process can be divided into five

different physical stages. The good agreement between the experiments and the numerical

model makes it possible to use the latter in design of similar systems.

Keywords: PCM, modelling, cold storage, dimensional analysis, insulation

2
1. INTRODUCTION

Phase-change materials (PCMs) are considered suitable for various industrial

applications due to their high energy density and their ability to provide a relatively constant

temperature during the solid-liquid phase-change process. These benefits can be useful in

situations where the desirable effect is preservation of a certain temperature range, like in

passive thermal management of electronic equipment [1] and realization of thermal comfort

in buildings [2], or where the objective is to accumulate heat for its further use, like in latent

heat thermal energy storage (LHTES) systems [3].

For LHTES applications, the PCM can be encapsulated in containers of different

shapes or orientations [4]. One of the most basic and common configurations are vertical

cylindrical or annular containers. These have been studied in the past, including three decades

ago by Sparrow and Broadbent [5], who investigated inward melting in a copper vertical

tube. Of interest were the effects of different wall temperatures and the initial temperature of

the solid on the melting rate, and it was found that natural convection in the melt and initial

subcooling affected the melting rate significantly. Sparrow et al. [6] continued to investigate

the same configuration by comparing pure with impure materials, observing the same typical

melting patterns and generalizing the experimental results. Menon et al. [7] studied the

heating and melting rates for a vertical tube filled with a paraffin wax, confirming that natural

convection affects the solid shape and melting rate. A quasi-steady model that utilizes an

effective thermal conductivity, dependent on the Rayleigh number, was developed and its

results allowed satisfactory prediction of the melting rate. Wu and Lacroix [8] explored

numerically the problem of melting in a vertical cylindrical capsule, heated by a constant

temperature from all its walls. A complex, body-fitted co-ordinate model, that can track the

solid-liquid interface and takes into account convection in the melt, was developed. They

showed that the highest heat transfer rate was from the bottom due to natural convection. The

3
same authors also studied numerically melting of ice in a vertical cylinder heated from its

side [9]. Jones et al. [10] investigated, both experimentally and numerically, melting in a

short vertical cylinder, heated from its side. The phase-change modelling was conducted

using the enthalpy-porosity method [11] that takes into account natural convection in the

liquid phase. The melting front location and shape were monitored and compared with the

numerical results, showing a good agreement. Consistent with the previous works, natural

convection was found to be the dominant heat transfer mode for a considerable part of the

melting process. More recently, Shmueli et al. [12] investigated a similar configuration, both

experimentally and numerically, focusing on the local heat transfer rates and melt fractions.

Similarly to cylinders, vertical annular enclosures were also studied extensively in the

past. Wang et al. [13] investigated numerically an enclosure with an inner isothermal tube.

The model included the effect of natural convection by using the temperature transforming

method with high values of effective viscosity [14]. Dimensionless correlations for the

Nusselt number and the total and latent energy stored were derived. The same authors then

expanded their investigation to a tilted annular enclosure [15], comparing the results with

experimental findings. Longeon et al. [16] also studied a vertical annular enclosure, heated

from its inner tube. The effect of natural convection was found to be significant, both

experimentally and numerically. Annular vertical enclosures with fins were also explored, for

radially-finned [17] and longitudinally-finned [18] storage units.

Cylindrical and annular enclosures were also investigated as a part of a complete

storage systems. For instance, Esen and Ayhan [19] studied numerically a solar LHTES unit

that can improve the performance of domestic heating systems with heat pumps. The PCM

was encapsulated in cylindrical containers, heated by a heat transfer fluid (HTF), which flows

parallel to their axis inside the shell. An axisymmetric model, based on heat conduction and

coupled to the energy equation of the HTF, was developed. Then, a parametric investigation

4
was conducted, for the effects of the HTF inlet temperature and flow rate, the number of

cylindrical containers and their radii, and the PCM type, in order to meet the required system

performance. Esen et al. [20] expanded that investigation for the case where the HTF flowed

in the tubes and the PCM was encapsulated in the shell around them, finding that the new

configuration allowed higher melting rates than the previous one. Optimized dimensions were

derived for each configuration and different PCM types. A complete seasonal analysis for the

system was conducted by a similar model, developed by Esen [21] and validated

experimentally vs. the results of Çomakli et al. [22]. In order to improve the performance, it

was suggested to reduce the length of the tubes, the storage tank height, and the cylindrical

containers’ wall thickness. A similar problem, with a crossflow between the HTF and the

cylinders, was solved analytically, under different simplifications, by Dubovsky et al. [23].

It is thus evident that LHTES in vertical cylindrical and annular enclosures, and in

particular their numerical modelling, is extensively addressed in the literature. However, the

current work deals with a different application, which, though related to heat storage, is less

frequent in the literature. Specifically, there are some applications where it is desired to keep

the “cold storage”, based on a certain pre-frozen material, at a low temperature for as long as

possible, rather than trying to heat it up and melt rapidly with a hot HTF, like in LHTES

charging. Also unlike LHTES, the heat in these applications is typically coming from the

ambient, by natural convection and radiation, which have considerably higher thermal

resistances than forced convection from an HTF. In this sense, this situation is similar to

PCM-based thermal protection in buildings [2], although a cylindrical geometry is less

common there. Moreover, unlike in LHTES systems, in buildings the PCM-based protection

is almost always combined with some other sort of insulation, e.g. like a PCM-impregnated

board [24,25], attached to a commonly built wall.

5
The application explored in this work is the use of insulated and transportable cold

storage packages that can carry products sensitive to high ambient temperature, e.g., raw fish

[26], ice cream [27], or bio-medical goods [28]. As stressed above, in such cases it is desired

to sustain, for as long as possible, the low temperature of the product. Obviously, this can be

done even without a PCM, with only a “conventional” insulation, although it is important to

note that the product itself can be frozen, and thus phase change from solid to liquid might be

encountered in the system. The use of various insulating materials for this purpose, without

any PCM, and their performance at different conditions have been studied in the past, both

experimentally and theoretically. For instance, Margeirsson et al. [26] studied the thermal

performance of different insulated packages with cold but unfrozen fish inside. The

temperatures were monitored and compared with results of a three-dimensional heat-

conduction-based numerical model, showing a good agreement between the two. Choi and

Burgess [29] developed a simplified model that predicts the thermal performance of cold

storage insulated packages of that sort, comparing it with experimental results for packages

with different dimensions under different environmental conditions. The tested product was

ice, assumed as a lumped heat capacity. Mittal and Parkin [30] investigated insulated ice-

cream containers with different sizes and insulating materials, and developed a three-

dimensional numerical model based on heat conduction, without considering the effect of

solid-liquid phase change. The predicted temperatures were within ±5% of the experimental

data. Zuritz and Sastry [31] developed an analytical model, based on a power series, for the

temperature variation in an insulated ice-cream container, exposed to periodic environmental

fluctuations, showing a good agreement with experimental results. The effects of the

insulation thickness and the heat transfer coefficient to the ambient were also investigated,

but not the process of solid-liquid phase change. Dolan et al. [32] created a finite-difference-

based one-dimensional model for heating of frozen food exposed to solar radiation, where the

6
solid-liquid phase change was modelled using the well-known effective heat capacity method

[33], reporting a reasonable agreement with the experimental results. It was found that the

temperature profile is highly dependent upon the thermal properties of the frozen food and its

initial temperature, while it is insensitive to the heat transfer coefficient at the exposed outer

surface. Moureh and Derens [34] used a three-dimensional finite-volume computational fluid

dynamics (CFD) software for temperature prediction in packaged frozen fish pallets. The

model showed a good agreement with the experimental results. However, the effect of the

solid-liquid phase change in the product was not considered.

Another possible approach to enhance the cold storage package performance is

replacing part of the package insulation with a solid PCM, which by its melting prevents heat

transfer to the product while allowing an almost constant low temperature at the inner part of

the package. Although this idea has been studied in the past, it is not very common in the

literature. For instance, Oró et al. [35] used this approach to improve the performance of

chilly bins that contained water or ice-cream. The thermal performance was tested with and

without PCM, finding that storage time can be increased by more than 300% by the use of

PCM. Also, a 2-D axisymmetric heat conduction-based numerical model was developed and

validated with the experiments. Oró et al. [27] studied the heating rate of a cold ice cream

container that was exposed to ambient temperature. It was shown that by surrounding the

container with a relatively small amount of PCM, the heating rate received by ice cream

decreased significantly. The experimental results validated a 3-D heat conduction model that

utilized the effective heat capacity method for modelling of solid-liquid phase change and

was then used for optimizing the PCM thickness at different walls of the container. Another

work on ice cream storage [36] showed that the use of PCM can sustain an almost constant

storage temperature, unlike conventional insulation, where the temperature rises more

rapidly. Laguerre et al. [37] investigated a container with cold products that were maintained

7
at a low temperature by using insulation and PCM, suggesting also a simplified empirical

model, found to give a good prediction until the PCM has melted. East and Smale [38]

developed a 3-D numerical model for heating of a cold storage package with PCM, using a

genetic algorithm to design a package with a minimal cost, while maintaining the required

thermal performance. This approach has also been utilized for an analysis under different

environmental conditions [39].

The use of different frozen gels as PCMs has also been investigated. Alasalvar and

Nesvadba [40] measured the latent heat capacity of different commercial cooling gels, then

investigating the thermal performance of ice and the gels by monitoring the temperature of

smoked salmon shipped at ambient temperature conditions. The temperature distribution was

also predicted by a numerical simulation. Stubbs et al. [41] developed a three-dimensional

numerical model for prediction of the temperature of chilled foodstuffs that are cooled by a

gel and insulated by expanded polystyrene. The model was validated against an experiment,

and the results showed that the temperature was sensitive to the food and gel positions. Elliott

and Halbert [42] studied experimentally insulated transit containers with frozen gel packs,

which were designed to keep the product at a specific temperature range. Optimal design was

found according to certain seasonal conditions.

Although all these works on cold storage packages show the effect of different

parameters at various conditions, the roles of PCM and “conventional” insulation in the

combined systems need a further investigation. For instance, since they both contribute to the

same purpose, it is not clear if there is an optimal insulation thickness that can ensure the

desired maximum melting time of the PCM, thus maximizing the positive effect of the latter.

Also, it has been shown above that melting in cylindrical and annular enclosures is strongly

affected by free convection [5,10,12], whereas in the works on PCM cold storage packages

this effect is usually not taken into account, and the only heat transfer mode included is

8
conduction. In the present work, numerical models that indirectly take convection into

account, while still being inexpensive computationally and reflecting a three-dimensional

insulation, are developed, tested, validated against experimental results, and applied to real

situations. In addition, an analytical model that allows revealing of the governing

dimensionless groups and the existence of an optimal insulation thickness is suggested.

2. EXPERIMENTAL

The analysis performed in this work is based on an extensive comparison with

experiments. For this purpose, two different packages were tested. It is of interest to compare

between the performances of the two, in order to gain a better understanding on this type of

systems. Each of these packages is designated to carry a product and keep it at a low

temperature for a long time, while being transported. The packages have external dimensions

of 32×25×25 cm and 50×50×50 cm, and are referred to as the “smaller package” and the

“larger package”, respectively. Each package contains an insulation, some type of PCM, and

a product that is supposed to be kept at a low temperature. The smaller package is comprised

from the following parts: (1) A fitting cardboard box with 4 mm thick walls; (2) A special

plastic container with a double wall and an inner cavity for the product. The double wall is

filled with a green aqueous solution of salts, termed “PCM A”. Its melting temperature is -10

°C, and its other properties are found in Table 1; (3) High-density polystyrene insulation,

comprised of two halves, a lower part and a lid. The two halves form a cylindrical cavity that

fits the container size. Figure 1a shows the open cardboard box, with the lower part of the

insulation and the container inside.

The larger package is comprised from the following parts: (1) A fitting cardboard box

with 4 mm thick walls; (2) A thin-walled plastic container for the product, Fig. 1b (left);

(3) Four arched bottles, Fig. 1b (right), partially filled with a purple aqueous solution of salts,

9
termed “PCM B”. Its melting temperature is -33 °C, and its other properties are found in

Table 1; (4) High-density polystyrene insulation, comprised of two halves, a lower part and a

lid. The two halves form a cylindrical cavity that fits the container and the arched bottles.

Figure 1c shows the open cardboard box, with the lower part of the insulation and the four

arched bottles arranged in a circle shape around the container. The exact dimensions and the

basic geometry of the two packages are shown in Fig. 2 by a cross section view.

The experimental methodology is designed to accomplish two objectives: (a) To test

the packages’ best thermal performance, and (b) To allow validation of numerical simulations

with the experimental results. In order to meet objective (a), it was decided to freeze the PCM

in a -70 °C deep-freeze refrigerator, which is the lowest possible operating temperature

suggested by the packages’ supplier. This guarantees the longest time duration possible for

sustaining the product at a low temperature. In order to meet objective (b), water (H2O) is

used as the product to be transported. Water should be a good choice to approximate product

behaviour, because the sensible and latent heat capacities of water are comparable to those of

the real product. So, for each package, its product container is filled with tap water, up to the

height of the PCM. The water and air temperatures inside the container, and the ambient

temperature, are measured with standard T-type thermocouples (TCs). These measurements

are then used for experimental validation of numerical simulations. The TCs are connected to

the container’s lid, and are equally spaced: their exact location is depicted, for smaller and

larger packages, in Fig. 2a and Fig. 2b, respectively. Also, the measurement error of the TCs,

as given by the supplier, is ±1 °C.

The experimental procedure is similar for both of the packages and consists of the

following stages: (1) The container and/or the arched bottles are held in a -70 °C deep-freeze

refrigerator, for more than 36 hours, which is long enough to ensure that the PCM and the

product in the container have completely solidified and reached equilibrium temperature;

10
(2) The container and/or the PCM bottles are sealed inside the polystyrene insulation and

inserted into the cardboard package; (3) The package is placed on a wooden table, while all

its other sides are directly exposed to ambient temperature; (4) The temperatures are

measured for every 60 seconds until the end of the process, assumed when the product

temperature is sufficiently close to the ambient temperature.

3. MODELING

The theoretical analysis includes the use of three different models: a simplified one-

dimensional analytical model and one- and two-dimensional numerical models, which are

implemented using the commercial software package ANSYS Fluent 13.

3.1 Analytical model

A simplified analytical model is developed to provide fast prediction and to reveal

different features of the problem, under the following assumptions:

(1) 1-D axisymmetric heat conduction takes place; (2) The process is quasi-steady; (3) The

solid PCM temperature remains at the melting temperature; (4) The material properties, heat

transfer coefficient to the ambient and the ambient temperature are constant. The general

schematics of the model are shown in Fig. 3.

The temperature distributions in the PCM and in the insulation, under assumptions (1)

and (2), have the following form:

T  r , t   A ln  r   B (1)

Tin  r , t   C ln  r   D (2)

where T is PCM temperature, Tin is insulation temperature, r is spatial radial coordinate and

A, B, C and D are constants to be determined.

The following boundary conditions are applied:

11
T  r  R  t  , t   Tm (3)

kin
dTin
dr r  Ro

 ho Tr  Ro  T  (4)

dT dT
kl  kin in (5)
dr r  Rp dr r  Rp

T  r  Rp   Tin  r  Rp  (6)

where R(t) is solid-liquid interface position, Tm is melting temperature, T∞ is ambient

temperature, ho is heat transfer coefficient to ambient, kl and kin are thermal conductivities of

liquid PCM and insulation, respectively, and Ro and Rp are the external radii of the insulation

and PCM, respectively.

By substituting Eqs. (1) and (2) into the boundary conditions, Eqs. (3-6), the

temperature distributions in the PCM and the insulation can be found:

T  Tm
T  r , t   Tm  ln  r / R  t   (7)
 kl kl 
 ln  Rp / R  t    ln  Rp / Ro   
 kin Ro ho 

T  Tm  kin 
Tin  r , t   ln  r / Ro     T (8)
 kin kin   Ro ho 
 ln  Rp / R  t    ln  Rp / Ro   
 kl Ro ho 

The Stefan condition [43] at the solid-liquid interface is used in order to find the

melting rate:

dR dT
L  kl (9)
dt dr r  Rt 

By substituting Eq. (7) into Eq. (9) we derive the following expression:

dR k T  Tm
L  l (10)
dt R t   kl kl 
 ln  Rp / R  t    ln  Rp / Ro   
 kin Ro ho 

This expression can be integrated:

12
Rt 
  kl kl   kl
t

  R  t  ln  R
 p / R  t   
 R h k  ln 
 p o 
R / R  R  t   dR   0 T  Tm  dt (11)
Rp   o o in    L

yielding the following transcendental equation, which allows calculation of the solid-liquid

interface position as a function of time:

 2kl  2 t
 2 ln  Ro / Rp   1  R  t   Rp   
kl 4kl
2 R  t  ln  Rp / R  t     T  T  dt
 L 0  m
2 2
(12)
 Ro ho kin 
The total melting time can be derived by substituting R(t=tmelt)=Ri, which is the

internal radius of the PCM:

L  2kl  
 2 l ln  Ro / Rp   1  Rp2  Ri2   2 Ri2 ln  Rp / Ri  
k
tmelt   (13)
4kl T  Tm   Ro ho kin  

3.2 Estimating the heat transfer coefficient

It is known that air heat transfer depends on the wall temperature, and is governed by

natural convection and thermal radiation. An effective heat transfer coefficient, which takes

into account both convection and radiation, can be defined as follows:

heff  hnc  hrad (14)

The natural convection heat transfer coefficient is estimated using the following

correlation developed for natural convection of air on a vertical wall [44]:

1/4
 T T 
hnc  1.42  w   (15)
 H in 
where Tw is the insulation outer wall temperature and Hin is the insulation height.

The radiation heat transfer coefficient is defined as follows:

 (Tw4  T4 )
hrad  (16)
Tw  T

where  is the insulation outer wall emissivity and  is the Stefan-Boltzmann constant.

13
For the analytical model, the heat transfer coefficient was assumed constant,

according to assumption (4), and was determined as an average value between initial and

final values. Thus, heff was found equal to 10 W/m2K, using a trial and error procedure, for

the case of the smaller package.

3.3 The analytical model vs. numerical model

The results of the analytical model are first compared to a numerical model, in order

to test the influence of the simplifying assumptions used for the analytical model. Hence, a

one-dimensional numerical analysis was conducted using the commercial CFD software

ANSYS Fluent. Two cases were considered, one where heff ≈10 W/m2K and the other where

heff was calculated at each time step according to Eqs. (14-16). The dimensions are those of

the smaller package, and the PCM and polystyrene properties are given in Tables 1 and 2,

respectively. Figure 4 shows the melt fraction of the PCM as a function of time using the

analytical model and these two cases of constant and variable heff. The results are quite

similar, while the melting rate of the analytical model is slightly higher. This can be explained

by the fact that sensible heating is not taken into account in the analytical model. The

analytical model can also predict the time-dependent temperature distribution in the

insulation and the PCM, as shown in Fig. 5. Two different instances are shown, at the

beginning of the process (initial) and at the end of the process, after all the PCM has melted

(final). The temperature distributions did not change significantly with time, and the outer

insulation temperature remained almost constant. As expected, the high temperature gradient

occurred in the insulation due to its low thermal conductivity.

3.4 Optimal insulation thickness

It is clear that both insulation and PCM play an important role in keeping product

temperature low for an extended period. If the external dimensions are kept constant, then

14
increasing thickness of PCM means decreasing insulation thickness, and vice versa. Thus, the

question is whether there exists an optimum thickness of insulation and PCM for fixed

package external dimensions, which gives a maximal time for the PCM to melt. Figure 6a

shows the melting time as a function of PCM radius as calculated from Eq. (13). Two limiting

cases are shown: where the PCM radius is equal to the inner radius, i.e. there is no PCM, and

where the PCM radius is equal to the outer radius, meaning that there is no insulation. It can

be seen that for the first option the melting time is of course zero because there is no PCM.

For the second option, the melting time is obviously not zero, but it turns out that it is also not

the highest value possible. Figure 6 shows that there is an optimal PCM radius that enables

maximizing the melting time. This optimum point occurs because of the trade-off between

the PCM large heat capacity (“thermal mass”) and the insulation low thermal conductivity.

The melting time according to the smaller package used in this work is marked by the

“square” symbol (current case). It appears that the PCM radius chosen in the design of the

packages is quite close to the theoretical optimal value. In order to generalize the problem,

for any geometry and PCM content, a dimensional analysis can be utilized.

3.5 Dimensional analysis

Equation (13) can be rendered into a dimensionless form as follows:

1  2  
FoSte    2k* ln  R* / R p*   1  R p2*  1  2ln  R p*   (17)
4  Bi  
based on the following dimensionless groups:

 l tmelt
- Fourier number: Fo 
Ri2

c p T  Tm 
- Stefan number: Ste 
L

- Dimensionless PCM radius: Rp*  Rp / Ri

15
ho Ro
- Biot number: Bi 
kl

- Ratio of the thermal conductivities: k*  kl / kin

- Ratio of the radii: R*  Ro / Ri

It is possible to find a value for Rp*, which provides maximum melting time, by

differentiating Eq. (17) and equating the result to zero:

d  FoSte  1  2  
   2k* ln  R* / Rp*   1  Rp2*  1  2ln  Rp*    0 (18)
dRp* 4  Bi  

The following transcendental equation allows finding the critical dimensionless PCM radius,

Rp*cr , which guarantees the maximal melting time:

 2 
Rp*2cr 1  k*   2k* ln  R* / Rp*cr    1  k* (19)
 Bi 

According to this analysis, the results of Fig. 6a have been rendered into

dimensionless form, showing the dimensionless time, FoSte, against the dimensionless PCM

radius, Rp*. One immediate conclusion from Eq. (19) is that the critical PCM radius, under

these assumptions, does not depend at all on the Ste number. This means that this analysis is

general for any PCM latent heat, melting temperature and ambient temperature. However, the

PCM thermal conductivity does affect the critical PCM radius value. Two dimensionless

groups that govern the PCM critical radius are: (1) the Biot number, Bi, and (2) the ratio of

PCM to insulation thermal conductivities, k*. In Fig. 6b, the maximal value of the

dimensionless melting time is also marked by a square. Figure 7a shows the effect of Bi

number on melting time. One can see that an increase in the Bi number shortens the melting

time, due to the increased heat transfer coefficient from ambient. This also changes the

position of the optimal point slightly, while a higher Bi number requires a thicker layer of

16
insulation in order to maximize package performance. It is important to note that although Bi

number is varied widely, the change in the dimensionless melting time is relatively small.

Figure 7b shows that when thermal conductivity of insulation decreases (increase in k*) the

melting time increases, which is expected. In comparison with Fig. 7a, it appears that melting

time is much more sensitive to k* than to Bi. This is similar to the results of Dolan et al. [32],

as their solution was considerably more sensitive to the frozen food thermal conductivity than

to the outer surface heat transfer coefficient. However, the optimal PCM radius does not

change significantly when any of these parameters is changed. This makes Eq. (19) very

useful for the design of industrial packages containing PCM.

3.6 Numerical model

The analytical model reveals the basic features of the problem, but its specific results

are of limited applicability because the melting process involves (1) sensible heating, (2)

temperature-dependent heat transfer coefficient, (3) three-dimensional heat transfer and (4)

natural convection in the PCM, product and air cavities. A full solution to the problem

requires solving the 3-D Navier-Stokes equations coupled with the energy equation, which is

time-consuming. Thus, in this study the numerical model is developed under the following

considerations. By observing the geometry of the packages, one can see that the container and

the PCM bottles are axisymmetric, while the polystyrene insulation is not. It is possible to

consider an approximate approach, which does not involve a three-dimensional solution,

where the model is entirely axisymmetric. An effective radius Reff of the insulation can be

defined, by equating the actual insulation cross-section to the insulation with an effective

radius Reff, as shown below:

 2Ro    Rp2   Reff2   Rp2


2
(20)

17
where Ro and Rp are the insulation outer and inner radii. The following relation can be thus

derived:

2
Reff  Ro (21)

In order to verify the accuracy of this approach, it is tested for three different cases:

the analytical and 1-D CFD models, both with an effective insulation radius according to Eq.

(21), and a two-dimensional CFD planar model. Figure 8a shows the computational domain

that is solved for the latter, where only 1/8 of the full domain needs to be solved due to

symmetry. The outer insulation wall boundary condition is determined according to Eq. (14).

Figure 8b shows the melt fraction as a function of time for the three cases. All cases show

similar results, which means that this approximation takes into account the non-axisymmetric

heat conduction, and can be used as an improvement of either the analytical model or the 1-D

CFD model. The actual effect of this improvement in comparison with the axisymmetric

model can be estimated against the results shown in Fig. 4. It can be seen from Fig. 8b that

the melting time for the improved model is increased by more than 10%, which is reasonable

because of the thicker insulation layer.

Another important feature that can change the results significantly is natural

convection in the melted PCM, product and air cavities. Natural convection in the melt of the

PCM is modelled by an effective thermal conductivity keff that is dependent on the Rayleigh

number, Raδ, defined based on the average molten layer thickness, δ [45]:

 gβ  Tw  Tm   3 
n

keff / k  cRa  c 
n
 (22)
 να 

where g is the gravitational acceleration,  is the volumetric expansion coefficient,  is the

kinematic viscosity, and  is the thermal diffusivity. According to [44], the constant c

depends on the radius and the height H of the PCM, and its value was estimated to be 0.22

18
and 0.24, for the smaller and larger packages, respectively; the value of the constant n is 0.25,

while this correlation holds for Raδ>2000. The molten layer average thickness is calculated

according to the definition of the PCM melt fraction MF for an annulus with a uniform

molten layer thickness:


π  Rp2  R p  δ  H 
2

 
MF   (23)
π  Rp2  Ri2  H

From Eq. (23) the following relation for the average molten layer thickness can be

derived:

δ  Rp  Rp2  MF  Rp2  Ri2  (24)

For the product (water) inside the container a similar relation is developed, taking into

account the cylindrical shape of the enclosed water:


δ  Ri 1  1  MF2  (25)

where MF2 is the product melt fraction. It is assumed that Eq. (22) holds also after the entire

PCM or product has melted, but the temperature difference is calculated with the average

liquid temperature instead of the melting temperature.

Natural convection can also occur in the air cavities. However, most of these cavities

are heated from above (conduction heat transfer), and those that are heated from the side have

relatively low height in comparison with their width. The only air cavity that is heated from

below is the one which is under the PCM and the product in the smaller package. According

to these considerations, only the latter cavity is modelled for natural convection by an

effective thermal conductivity:

 gβ  Tw 2  Tw1  Z 3 
n

keff / kair  cRa  c 


n
Z  (26)
 να 

19
where Z is the height of the air cavity below the PCM, and Tw2 and Tw1 are the average

temperatures of the lower and upper walls of the cavity. The values of the constants c and n

are given according to [44]: 0.059 and 0.4, for 1700<RaZ<7000, 0.212 and 0.25, for

7000< RaZ<3.2×105, and 0.061 and 1/3 for RaZ>3.2×105, respectively.  is calculated with the

ideal gas law assumption as the inverse of the average between wall temperatures in Kelvin.

By using these simplifications, a two-dimensional axisymmetric heat conduction

problem is solved by ANSYS Fluent. It involves only the energy equation while the solid-

liquid phase change is modelled by the enthalpy method, so no explicit tracking of the solid-

liquid interface is necessary [46]. The following form of the energy equation is solved:

H
     k T  (27)
t

where H is specific enthalpy, defined as follows:

H  h  H (28)

where h and H are specific sensible and latent enthalpies, respectively.

The latent heat is dependent on temperature according to the following relation:

L T  Tl

H T    fL Tl  T  Ts (29)
0 T  Ts

where f is the melt fraction at each cell, and Tl and Ts are the liquidus and solidus

temperatures, respectively. The transient formulation is first-order implicit and the spatial

discretization for the energy equation is done with a first-order upwind scheme. The residual

at each time step for the energy equation was determined to be 10-9.

The model includes the insulation with an effective radius, the PCM, the product

(water) inside, the air cavities below and above the PCM and also the wooden table on which

the packages were placed. Heating from ambient includes radiation and natural convection

20
that are estimated using an effective heat transfer coefficient, see Eq. (14). The natural

convection heat transfer coefficient from the side of the insulation is estimated by Eq. (15),

where the characteristic scale is the package height. Heating by natural convection from the

top of the package and the bottom of the wood (table) is estimated according to the following

correlation for air [44]:

1/4
 T T 
hnc  C1  w   (30)
 D /4
 eff 

where Deff is the package effective diameter and C1 is a constant, and its values are 0.59 and

1.32 for heating from above and from below, respectively.

The volumetric expansion coefficient of the product (water) and PCM A is taken as an

average value between its values at the melting and ambient temperatures. Its value is

assumed to be =2×10-4 1/K, except for PCM B in the larger box, where the temperature

difference is higher, and it is taken as  =3×10-4 1/K. All the other properties are constant,

except for the PCMs and the product specific heat and thermal conductivity, which account

for melting and changes according to the phase distribution in each computational cell. The

phase change is assumed to occur in a narrow temperature range, 0.1 C, where both thermal

conductivity and specific heat change linearly from the solid to the liquid. The PCMs and

product (water) properties, as estimated by data given from the manufacturer, are shown in

Table 1, and the properties of the polystyrene insulation and wood are shown in Table 2.

It is important to stress that thermal resistances of the thin cardboard and plastic

layers, and also the different contact resistances, are considered negligible in comparison with

the large thermal resistance of the thick polystyrene insulation.

21
4. COMPARISON BETWEEN EXPERIMENTS AND MODEL PREDICTIONS

4.1 Verification of the numerical results

It is important to first verify that the numerical solution does not depend on the grid or

the time-step sizes. Thus, for each package, two different grids and time-steps were tested.

For both packages, grids with square cells and sizes of 0.25 mm and 0.125 mm were used.

This corresponds to 7590 and 36360 cells, for the smaller package, and 22000 and 88000

cells, for the larger package, respectively. Time-steps of 1 and 0.5 seconds were tested for the

smaller package, and those of 5 and 2.5 seconds for the larger package. Figures 9a and 9b

show the temperature evolution of TC 1, with the different grids and time-steps, for the

smaller and larger packages, respectively. It is easy to see that for both packages the different

curves are almost identical, and that the solution is independent of the grid and time-step

sizes.

4.2 Validation of the model

The numerical model, for the smaller package, is validated against the experiment.

Figure 10 shows a comparison between the measured and predicted temperatures at the same

location. The entire process may be divided into 5 stages: initial sensible heating up to the

melting temperature of PCM A (-10°C), melting of the PCM A at an almost constant

temperature, sensible heating up to the product (water) melting temperature (0°C), melting of

the product at an almost constant temperature, and finally sensible heating to ambient

temperature. These trends are also captured by the simulation for all thermocouples. Also, up

until the end of the fourth stage, both the experimental and numerical results show that TCs

1-4 have almost identical temperatures, and there is a good agreement between these results.

Then, for both the experiment and simulation, the temperatures of the lower TCs become

22
higher than the upper ones. However, the melting period of the fourth stage is slightly longer

for the experiment than the simulation, which also affects the temperature prediction of the

fifth stage. This discrepancy can be explained by the approximate prediction of natural

convection and also by the uncertainties we may have in materials properties. TCs 5 and 6

show reasonable agreement at the first stage, afterwards the measured TCs 5 and 6 are higher

and lower than the simulation, respectively. At the beginning of the fifth stage, there is once

again good agreement. According to the simulation, the location of TC 5 is exactly at the

edge of the frozen water, but in the experiment it is possible that this is influenced by the air

temperature that is above the frozen water. For TC 6, some of the discrepancy can be related

to the assumption that natural convection at the upper air cavities can be neglected. Some of

the observed trends can be found in other works from the literature. For instance, Oró et al.

[35] showed, both numerically and experimentally, that the product temperature, for water

and also ice cream, is rising until reaching the PCM melting temperature. Then, the

temperature rises very slowly for a long period of time, and starts rising faster again after the

melting process is over. This is similar to stages 1-3 in Fig. 10. However, in their study, the

part where the product itself is melting was not studied, so stages 4 and 5 cannot be

compared. On the other hand, Alasalvar and Nesvadba [40] showed results which are similar

to stages 3-5 in Fig. 10. In their work, there is a temperature rising up to the melting

temperature, a melting period with an almost constant temperature, and sensible heating,

where the slope is continuously decreasing due to the decrease in the temperature difference

between the product and the ambient. It is important to note that in their study, the product

was not frozen, hence an immediate rise of the temperature directly to the ambient

temperature was allowed after the PCM melting period.

The simulation allows us investigating different features of the problem that cannot be

estimated easily from the experiments. For instance, Fig. 11a shows the melt fraction as a

23
function of time for both PCM A and product (water). The melt fraction also gives a clear

indication on the onset of each one of the five stages observed in Fig. 10. The heat transfer

rate from ambient to the package can also be predicted as shown in Fig. 11b. Initially, the

outer package temperature is exactly the same as the ambient temperature and the heat

transfer rate is zero. Very quickly, however, the outer wall temperature drops due to the inner,

much colder, PCM A, and the heat transfer rate increases up to almost 8 W. Then, the heat

transfer rate starts to decrease monotonically, due to the increase in the outer wall

temperature. It seems that the decrease in the heat transfer rate, at this stage, is almost linear.

Once the melting of PCM A starts, the heating rate stays almost constant with only a very

slight decrease. Similar trends are observed for the next sensible heating and phase change

stages. The last sensible heating stage is longer than the other stages, and continues until

heating rate decreases exponentially to zero.

Figure 12 shows the temperature profile evolution at the different instances, for the

smaller package. Fig 12a presents the radial temperature profile at the cross-section plane in

the middle of the PCM layer. Four instances are presented, namely 1, 30, 60, and 90 hours

from the start of the simulation, and each of these instances presents a different stage of the

melting process. Generally, for all the instances, the radial temperature profile in PCM A and

the product is almost constant, whereas a large temperature gradient is found in the

insulation. This is expected because the thermal conductivity of the insulation is much lower

than those of the PCM or product (water). Obviously, the temperatures are rising as the entire

package heats up, and at times where the PCM (30 hr) or the product (60 hr) are phase

changing, their temperature is almost equal to the melting temperature. The temperature at the

outer surface of the package is always higher than 20 ºC and it is rising only by a few

centigrades through the entire process. Figure 12b shows the axial temperature profile at the

axis of symmetry of the package, and includes the wood (table), lower insulation, lower air

24
cavity, product, upper air cavity, and upper insulation. The same four instances, as in Fig.

12a, are presented. Initially, after 1 hour, the temperature gradients are quite significant. The

largest gradients are in the upper (stagnant) and lower air cavities, and in the upper and lower

parts of the insulation. The gradients in the wood and product are much lower, where for the

latter they are almost zero. This is quite expected due to the low thermal conductivity of the

air and insulation in comparison with the product (water) and wood. At this stage, the effect

of differences between each part is pronounced, and rather abrupt changes in the gradient are

visible. For the later stages of the melting, the gradients decrease significantly, and the

differences between the different parts are less pronounced. Still, the largest gradients are in

the insulation and air cavities, and the product is almost isothermal.

Figure 13 presents a comparison between the measured and predicted temperatures

for the larger package. Similar to the smaller package, the entire heating process comprises of

5 stages, which is also captured by the simulation. Thermocouples 1-6 in the experiment

show almost identical temperatures throughout the entire process, and this is also true for TCs

1-5 in the simulation. Only TC 6 in the simulation shows different values, this can be

explained by the fact that effects of air mixing by natural convection cannot be ignored. One

can see that the temperature in the first stage of the sensible heating is accurately predicted.

However, the predicted melting period at the second stage is longer than that measured. This

difference can be associated with the uncertainty in the PCM properties and the approximate

approach followed in the prediction of natural convection, which is more significant here in

comparison with the smaller package. One may note that here the melting process is much

longer than in the smaller package. This trend, that for larger cylindrical storage containers

the melting time increases significantly, was also observed in numerical simulations by Esen

and Ayhan [19]. Also, the long duration at an almost constant temperature of the product is

similar to that observed by Leducq et al. [36], where the advantage of using PCM instead of

25
only standard insulation is shown quite clearly.

Figure 14a shows the progression with time of the melt fraction of PCM B and the

product (water). One may note that the slope of the PCM B curve is higher than the product

curve. This is because the product is found at a more inner part of the package, and thus its

melting rate is lower. Similar to Fig. 11a, these curves show approximately the onset and end

for each of the five stages. One can also see that the last sensible heating (fifth) stage is quite

long, covering half of the total heating time. Figure 14b trends are very similar to Fig. 12b,

although the heating rates are higher. Also, the heating rate is almost constant at the phase

change stages, where for the smaller box it decreased at a higher rate.

Figure 15 shows the temperature profile evolution at the different instances, for the

larger package. Similar to Fig. 12a, Fig 15a shows the radial temperature profile at the cross-

section plane in the middle of the PCM layer. Five instances are shown, namely 1, 60, 120,

180, and 300 hours, where each one represents a different stage of the melting. Similar to

Fig. 12a, the largest gradients are found in the insulation, while the product and PCM are

almost isothermal. However, after 180 hours, at the 4th stage of the melting, there is a small

gradient between the two, and it seems that the PCM starts to heat up, while the product (ice)

is still undergoing the phase change. The outer temperature of the insulation is, always,

almost equal to the ambient temperature, and its rise with time is much slower in comparison

with the smaller package. Figure 15b is similar to Fig. 12b and shows the axial temperature

profile at the axis of symmetry of the larger package. It includes the wood (table), lower

insulation, product, upper stagnant air cavity, and upper insulation. Similar trends in

comparison with Fig. 12b are shown here: the largest temperature gradients are found in the

insulation and air cavity, a much lower gradient in the wood layer, and the product is almost

isothermal. These gradients decrease with time, but even after 180 hours, it is still easy to see

the differences between the various layers. Interesting trends are observed for the outer

26
surface temperatures at the lower and upper insulation layers. After 1 hour, the outer surface

temperatures are almost equal to the ambient temperature; as time passes, the temperatures

are slightly lowered, and then start to rise again slowly. This trend has not been seen for the

smaller package in Fig. 12b, where the outer surface temperatures were rising monotonically.

The reason for this difference lies in the transient nature of the heating process. Initially, the

PCM, the product and the air cavities are at a low temperature of -70 ºC, while the insulation

and wood are at the ambient temperature. After some time, the insulation is cooled due to the

lower temperatures at the inner part of the package. It seems that for the larger package, even

after 1 hour, the upper and lower outer surface temperatures are still unaffected by the lower

temperatures at the inner part of the package. This difference is related to the thicker layer of

insulation in comparison with the smaller package.

5. CONCLUSION

In this paper the problem of a cold storage package that includes insulation and PCM

was investigated experimentally and theoretically. A simplified analytical model was

developed in order to provide more insights to the problem. The model revealed the

dimensionless groups that govern the problem, and also showed that there is an optimal ratio

of insulation/PCM thicknesses that maximize the melting time. Also, the melting time is

highly dependent on the ratio between the PCM and insulation thermal conductivities, but not

very sensitive to Biot number. In addition, it was shown that the optimal insulation and PCM

thicknesses remain almost unchanged by variation of these two dimensionless groups.

For a more exact prediction of the experimental results, a two-dimensional

axisymmetric numerical model was developed. The model is then validated against

experiments, which include two packages, with different dimensions and PCMs. The

numerically predicted temperatures were compared with the experimental results, showing a

27
reasonable agreement and very similar physical trends. In particular, five different stages of

the process were clearly established both experimentally and numerically.

The proposed model allows relatively fast computation, while taking into account

complex physical phenomena involved. It can be used for design and optimization of cold

storage packages with different dimensions and materials, under various conditions and

constraints.

ACKNOWLEDGEMENT

This project has received funding from the Royal Society of New Zealand and the

European Commission Seventh Framework Programme (FP/2007-2013) under Grant

agreement No. PIRSES-GA-2013-610692 (INNOSTORAGE).

This publication is also part of a project that has received funding from the European

Union’s Horizon 2020 research and innovation program under grant agreement No. 657466

(INPATH-TES).

We also want to acknowledge Dangerous Goods Management Ltd. for providing the

packages.

28
REFERENCES

[1] Sahoo, S.K., Das, M.K., and Rath, P., “Application of TCE-PCM based heat sinks for
cooling of electronic components: A review,” Renew. Sust. Energ. Rev., 59, pp. 550–582,
(2016).
[2] Kuznik, F., David, D., Johannes, K., and Roux, J.J., “A review on phase change materials
integrated in building walls,” Renew. Sust. Energ. Rev., 15, pp. 379-391, (2011).
[3] Zalba, B., Marín, 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.,
23, pp. 251-283, (2003).
[4] Dhaidan, N.S., and Khodadadi, J.M., “Melting and convection of phase change materials
in different shape containers: A review,” Renew. Sust. Energ. Rev., 43, pp. 449–477,
(2015).
[5] Sparrow, E.M., and Broadbent, J.A., “Inward melting in a vertical tube which allows free
expansion of the phase-change medium,” J. Heat Transf., 104, pp. 309-315, (1982).
[6] Sparrow, E.M., Gurtcheff, G.A., and Myrum, T.A., “Correlation of melting results for
both pure substances and impure substances,” J. Heat Transf., 108, pp. 649-653, (1986).
[7] Menon, A.S., Weber, M.E., and Mujumdar, A.S., “The dynamics of energy storage for
paraffin wax in cylindrical containers,” The Canadian Journal of Chemical Engineering,
61, pp. 647-653, (1983).
[8] Wu, Y.K., and Lacroix, M., “Melting of a PCM inside a vertical cylindrical capsule,” Int.
J. Numer. Meth. Fl., 20, pp. 559–572, (1995).
[9] Wu, Y.K., and Lacroix, M., “Analysis of natural convection melting of a vertical ice
cylinder involving density anomaly,” Int. J. Numer. Method. H., 3, pp. 445–456, (1993).
[10] Jones, B.J., Sun, D., Krishnan, S., and Garimella, S.V., “Experimental and numerical
study of melting in a cylinder,” Int. J. Heat Mass Tran., 49, pp. 2724–2738, (2006).
[11] Voller, V.R., and Prakash, C., “A fixed grid numerical modeling methodology for
convection-diffusion mushy region phase-change problems,” Int. J. Heat Mass Tran., 30,
pp. 1709-1719, (1987).
[12] Shmueli, H., Ziskind, G., and Letan, R., “Melting in a vertical cylindrical tube:
numerical investigation and comparison with experiments,” Int. J. Heat Mass Tran., 53,
pp. 4082–4091, (2010).
[13] Wang, S., Faghri, A., and Bergman, T.L., “Melting in cylindrical enclosures: Numerical
modeling and heat transfer correlations,” Numer. Heat Tr. A-Appl., 61, pp. 837-859,
(2012).
[14] Cao, Y., and Faghri, A., “A numerical analysis of phase-change problems including
natural convection,” J. Heat Transf., 112, pp. 812-816, (1990).
[15] Sharifi, N., Robak, C.W., Bergman, T.L., and Faghri, A., “Three-dimensional PCM
melting in a vertical cylindrical enclosure including the effects of tilting,” Int. J. Heat
Mass Tran., 65, pp. 798-806, (2013).

29
[16] Longeon, M., Soupart, A., Fourmigué, J.F., Bruch, A., and P., Marty, “Experimental and
numerical study of annular PCM storage in the presence of natural convection,” Appl.
Energ., 112, pp. 175-184, (2013).
[17] Kozak, Y., Rozenfeld, T., and Ziskind, G., “Close-contact melting in vertical annular
enclosures with a non-isothermal base: theoretical modeling and application to thermal
storage,” Int. J. Heat Mass Tran., 72, pp. 114–127, (2014).
[18] Al-jandal, S.S., and Sayigh, A.A.M., “Thermal performance characteristics of STC
system with phase change storage,” Renew. Energ., 5, pp. 390–399, (1994).
[19] Esen, M., and Ayhan, E.M., “Development of a model compatible with solar assisted
cylindrical energy storage tank and variation of stored energy with time for different
phase change materials,” Energy Convers. Mgmt, 51, pp. 357-366, (1996).
[20] Esen, M., Durmuş, A., and Durmuş, A., “Geometric design of solar-aided latent heat
store depending on various parameters and phase change materials,” Sol. Energy, 62, pp.
19-28, (1998).
[21] Esen, M., “Thermal performance of a solar-aided latent heat store used for space heating
by heat pump,” Sol. Energy, 69, pp. 15-25, (2000).
[22] Çomakli, Ö., Kaygusuz, K., and Ayhan T., “Solar-assisted heat pump and energy storage
for residential heating,” Sol. Energy, 51, pp. 357-366, (1993).
[23] Dubovsky, V., Ziskind, G., and Letan, R., “Analytical model of a PCM-air heat
exchanger,” Appl. Therm. Eng., 31, pp. 3453-3462, (2011).
[24] Sari, A., Karaipekli, A., and Kaygusuz, K., “Capric acid and stearic acid mixture
impregnated with gypsum wallboard for low-temperature latent heat thermal energy
storage,” Int. J. Energy Res., 32, pp. 154-160, (2008).
[25] Kuznik, F., Virgone, J., and Noel, J., “Optimization of a phase change material wallboard
for building use,” Appl. Therm. Eng., 28, pp. 1291-1298, (2008).
[26] Margeirsson, B., Gospavic, R., Pálsson, H., Arason, S., and Popov, V., “Experimental
and numerical modelling comparison of thermal performance of expanded polystyrene
and corrugated plastic packaging for fresh fish,” Int. J. Refrig., 34, pp. 573-585, (2011).
[27] Oró, E., de Gracia, A., and Cabeza, L.F., “Active phase change material package for
thermal protection of ice cream containers,” Int. J. Refrig., 36, pp. 102-109, (2013).
[28] Mondeig, D., Rajabalee F., Laprie A., Oonk H.A.J., Calvet T., and Cuevas-Diarte M.A.,
“Protection of temperature sensitive biomedical products using molecular alloys as phase
change materials,” Transfus. Apher. Sci., 28, pp. 143-148, (2003).
[29] Choi, S.J. and Burgess, G., “Practical mathematical model to predict the performance of
insulating packages,” Packag. Technol. Sci., 20, pp. 369-380, (2007).
[30] Mittal, G.S. and Parkin, K.L., “Transportation of frozen food in insulated containers –
theoretical and experimental results,” J. Food Process Eng., 9, pp. 81-92, (1986).

30
[31] Zuritz, A.C. and Sastry, S.K., “Effect of packaging materials on temperature fluctuations
in frozen foods: Mathematical model and experimental studies,” J. Food Sci., 51, pp.
1050-1056, (1986).
[32] Dolan, K.D., Singh, R.P., and Heldman, D.R., “Prediction of temperature in frozen foods
exposed to solar radiation,” J. Food Process Pres., 11, pp. 135-158, (1987).
[33] Farid, M.M., “A new approach in the calculation of heat transfer with phase change,” 9th
International Congress on Energy and Environment, Miami, USA, December, (1989).
[34] Moureh, J. and Derens, E., “Numerical modelling of temperature increase in frozen
food packaged in pallets in the distribution chain,” Int. J. Refrig., 23, pp. 540-552, (2000).
[35] Oró, E., Cabeza, L.F., and Farid, M.M., “Experimental and numerical analysis of a chilly
bin incorporating phase change material,” Appl. Therm. Eng., 58, pp. 61-67, (2013).
[36] Leducq, D., NDoye, F.T., and Alvarez, G., “Phase change material for the thermal
protection of ice cream during storage and transportation,” Int. J. Refrig., 52, pp. 133-139,
(2015).
[37] Laguerre, O., Ben Aissa, M.F., and Flick, D., “Methodology of temperature prediction in
an insulated container equipped with PCM,” Int. J. Refrig., 31, pp. 1063-1072, (2008).
[38] East, A. and Smale, N., “Combining a hybrid genetic algorithm and a heat transfer model
to optimise an insulated box for use in the transport of perishables,” Vaccine, 26, pp.
1322-1334, (2008).
[39] East, A., Smale, N., and Kang, S., “A method for quantitative risk assessment of
temperature control in insulated boxes,” Int. J. Refrig., 32, pp. 1505-1513, (2009).
[40] Alasalvar, C. and Nesvadba, P., “Time/temperature profiles of smoked salmon packaged
with cooling gel and shipped at ambient temperature,” J. Food Sci., 60, pp. 619-621,
(1995).
[41] Stubbs, D.M., Pulko, S.H., and Wilkinson, A.J., “Wrapping strategies for temperature
control of chilled foodstuffs during transport,” T. I. Meas. Control, 26, pp. 69-80, (2004).
[42] Elliott, M.A. and Halbert, G.W., “Maintaining the cold chain shipping environment for
Phase I clinical trial distribution,” Int. J. Pharm., 299, pp. 49-54, (2005).
[43] Alexiades, V. and Solomon, A.D., “Mathematical modeling of melting and freezing
processes,” Taylor & Francis, (1993).
[44] Holman, J.P., “Heat transfer,” McGraw-Hill Companies, (2010).
[45] Farid, M., Kim, Y., Honda, T., and A., Kanzawa, “The role of natural convection during
melting and solidification of PCM in a vertical cylinder,” Chem. Eng. Commun., 84, pp.
43-60, (1989).
[46] Shamsundar, N. and Sparrow, E.M., “Analysis of multidimensional conduction phase
change via the enthalpy model,” J. Heat Transf., 97, pp. 333-340, (1975).

31
Nomenclature
cp specific heat, kJ/(kg K)
D diameter, m
f melt fraction
g gravitational acceleration, m/s2
h heat transfer coefficient, W/(m2 K)
h specific sensible enthalpy, J/kg
H height, m
H specific enthalpy, J/kg
k thermal conductivity, W/(m K)
L latent heat, kJ/kg
r radial coordinate, m
R radius, m
t time, s
T temperature, °C
Z air cavity height, m
Greek letters
 thermal diffusivity, m2/s
β volumetric expansion coefficient, 1/K
 molten layer thickness, m
specific latent enthalpy in a cell, J/kg
µ dynamic viscosity, kg/(m·s)
 kinematic viscosity, m2/s
 density, kg/m3
Subscripts
eff effective
i inner
in insulation
l liquid
m melting
o outer
s solid
w wall

32
Table 1. Properties of the product (water) and aqueous solutions of salts used in this study.

Property Unit Product PCM A PCM B


(water) Smaller package Larger package
Melting temperature, ºC 0 -10 -33
Tm
Latent heat, J/kg 336000 244000 215000
L
Specific heat solid, J/(kg K) 2030 2030 2030
cp,s
Specific heat liquid, J/(kg K) 4180 3500 3500
cp,l
Density, kg/m3 998 1100 1174

Viscosity, kg/(m s) Temperature Temperature Temperature
 dependent dependent dependent
Thermal conductivity solid, W/(m K) 2.2 2.2 2.2
ks
Thermal conductivity liquid, W/(m K) 0.6 0.6 0.6
kl

Table 2. Properties of the other different materials used in this study.

Property Unit Polystyrene Wood


Specific heat J/(kg K) 1500 2310
Density kg/m 3
30 700
Thermal conductivity W/(m K) 0.045 0.173

33
Figure captions

Figure 1. Experimental set-up:


a. the smaller package and its different components.
b. the container and PCM arched bottle, for the larger package.
c. the inside of the larger package.

Figure 2. Cross-section view with the dimensions and the different sections of:
a. the smaller package.
b. the larger package.

Figure 3. Schematics of the analytical model, grey, blue and deep blue colors denote the

insulation, liquid PCM and solid PCM, respectively.

Figure 4. Comparison between the analytic model and 1-D Fluent simulations.

Figure 5. Time-dependent temperature distribution in the insulation and the PCM. Two
different instances are shown, at the beginning of the process (initial) and at the end of the
process, after all the PCM has melted (final).

Figure 6. Optimal thickness analysis:


a. The effect of the PCM radius on the melting time
b. The effect of the dimensionless PCM radius on SteFo.

Figure 7. The effect of:


a. the dimensionless PCM thickness
b. the Biot number on SteFo.

Figure 8. The 2-D planar model:


a. Schematics
b. Melt fraction for the different cases

Figure 9. Grid and time-step independence test for:


a. The smaller package
b. The larger package

Figure 10. Comparison between the experimental and numerical results, for the smaller
package.

Figure 11. Numerical results for the smaller package:


a. PCM A and the product (ice) melt fraction as a function of time

34
b. Heat transfer rate from the ambient

Figure 12. Temperature profiles for the smaller package:


a. Radial direction at the cross-section plane in the middle of the PCM layer
b. Axial direction at the axis of symmetry

Figure 13. Comparison between the experimental and numerical results, for the larger

package.

Figure 14. Numerical results for the larger package:


a. PCM B and the product (ice) melt fraction as a function of time
b. Heat transfer rate from the ambient

Figure 15. Temperature profiles for the larger package:


a. Radial direction at the cross-section plane in the middle of the PCM layer
b. Axial direction at the axis of symmetry

35
a.

b.

c.

Figure 1.

36
a.

b.

Figure 2.

37
Figure 3.

38
1

0.8
Melt fraction

0.6

0.4
1-D analytic - h=10
1-D Fluent - h=10
0.2
1-D Fluent - h=Rad+NC

0
0 10 20 30 40 50 60
Time, hr

Figure 4.

39
25
initial
20
final
15
Temperature, °C

10
5
0
-5
-10
-15
0.0375 0.0575 0.0775 0.0975 0.1175
Radius, m

Figure 5.

40
60
current case
50

40
Melting time, hr

30

20

10

0
0.0375 0.0575 0.0775 0.0975 0.1175
Rp , m

a.

12
current case

9
FoSte

0
1 1.5 2 2.5 3 3.5
Rp*
b.

Figure 6.

41
15
Bi=1
Bi=2
12
Bi=4
9 max
FoSte

0
1 1.5 2 2.5 3 3.5
Rp*
a.

25
k*=7
k*=13
20
k*=26
max
15
FoSte

10

0
1 1.5 2 2.5 3 3.5
Rp*

b.

Figure 7.

42
a.

0.8
Melt fraction

0.6

0.4

planar - Fluent
0.2
1-D axisymmetric Fluent - effective radius
1-D axisymmetric analytic - effective radius
0
0 10 20 30 40 50 60 70 80
Time, hr
b.

Figure 8.

43
30
20
10
Temperature, °C

0
-10
-20
-30
-40 0.25x0.25 mm, Δt=1 s
-50 0.25x0.25 mm, Δt=0.5 s
-60 0.125x0.125 mm, Δt=0.5 s
-70
0 30 60 90 120 150
Time, hr
a.

30
20
10
Temperature, °C

0
-10
-20
-30
0.25x0.25 mm, Δt=5 s
-40
0.25x0.25 mm, Δt=2.5 s
-50
0.125x0.125 mm, Δt=2.5 s
-60
-70
0 60 120 180 240 300 360 420 480
Time, hr
b.

Figure 9.

44
30
20
10
Temperature, °C

0
-10
-20
-30 TC 1 TC 2
TC 3 TC 4
-40 TC 5 TC 6
TC 7 TC 1s
-50 TC 2s TC 3s
TC 4s TC 5s
-60 TC 6s

-70
0 30 60 90 120 150
Time, hr

Figure 10.

45
1
PCM
Water
0.8
Melt fraction

0.6

0.4

0.2

0
0 30 60 90 120 150
Time, hr
a.

8
Heat transfer rate, W

0
0 30 60 90 120 150
Time, hr

b.

Figure 11.

46
30
20
10
Temperature, °C

0
-10
-20
-30
-40 1 hr
30 hr
-50
60 hr
-60 90 hr
-70
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Radial position, m
a.

30
20
10
Temperature, °C

0
-10
-20
-30
-40 1 hr
30 hr
-50
60 hr
-60 90 hr
-70
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Axial position, m
b.

Figure 12.

47
30
20
10
Temperature, °C

0
-10
-20 TC 1 TC 2
TC 3 TC 4
-30 TC 5 TC 6
TC 7 TC 8
-40 TC 1s TC 2s
TC 3s TC 4s
-50 TC 5s TC 6s

-60
-70
0 60 120 180 240 300 360 420 480
Time, hr

Figure 13.

48
1

0.8
Melt fraction

0.6

0.4

0.2 PCM
Water
0
0 60 120 180 240 300 360 420 480
Time, hr
a.

15
Heat transfer rate, W

12

0
0 60 120 180 240 300 360 420 480
Time, hr

b.

Figure 14.

49
30
20
10
Temperature, °C

0
-10
-20
-30
1 hr
-40 60 hr
-50 120 hr
180 hr
-60 300 hr
-70
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial position, m
a.

30
20
10
Temperature, °C

0
-10
-20
-30
1 hr
-40 60 hr
-50 120 hr
180 hr
-60 300 hr
-70
0 0.1 0.2 0.3 0.4 0.5
Axial position, m
b.

Figure 15.

50
Highlights

 Cold storage packages that utilize both PCM and regular insulation are studied
 An analytical model reveals the dimensionless groups that govern the problem
 Optimal configurations are determined in dimensional and dimensionless form
 A numerical model is developed to account for 3-D effects and natural convection
 Modeling approach is validated experimentally with a reasonable agreement

51

You might also like