PCM Book 1 PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 97

SPRINGER BRIEFS IN APPLIED SCIENCES AND TECHNOLOGY

THERMAL ENGINEERING AND APPLIED SCIENCE

Amy S. Fleischer

Thermal Energy
Storage Using
Phase Change
Materials
Fundamentals and
Applications
123
SpringerBriefs in Applied Sciences
and Technology

Thermal Engineering and Applied Science

Series editor
Francis A. Kulacki, Minneapolis, MN, USA
More information about this series at http://www.springer.com/series/10305
Amy S. Fleischer

Thermal Energy Storage


Using Phase Change
Materials
Fundamentals and Applications

13
Amy S. Fleischer
Department of Mechanical Engineering
Villanova University
Villanova, PA
USA

ISSN  2191-530X ISSN  2191-5318  (electronic)


SpringerBriefs in Applied Sciences and Technology
ISSN  2193-2530 ISSN  2193-2549  (electronic)
SpringerBriefs in Thermal Engineering and Applied Science
ISBN 978-3-319-20921-0 ISBN 978-3-319-20922-7  (eBook)
DOI 10.1007/978-3-319-20922-7

Library of Congress Control Number: 2015942817

Springer Cham Heidelberg New York Dordrecht London


© The Author(s) 2015
This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or
dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt
from the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained
herein or for any errors or omissions that may have been made.

Printed on acid-free paper

Springer International Publishing AG Switzerland is part of Springer Science+Business Media


(www.springer.com)
Preface

Phase Change Materials are being used for energy storage and thermal abatement
in a wide range of applications. These applications cover a wide range of sizes:
from small portable electronics to large-scale concentrating solar plants; and a
wide range of temperatures: from the −40 °C of space-based application to the
500 °C and up of solar energy applications.
In order to properly work with these fascinating materials, it is necessary to
understand their fundamental physical behavior, their thermophysical properties,
and the challenges inherent in working with them.
This monograph is intended to provide a comprehensive overview of phase
change materials and the current state of the art in their design and application.

Villanova, PA, USA Amy S. Fleischer

v
Acknowledgments

I wish to thank many of the students who have worked with me over the years on
various aspects of phase change materials research, particularly Ronald Warzoha,
Omar Sanusi, Kireeti Chintakrinda, Brian McManus, Rebecca Weigand, Kieran
Hess, Ryan Ehid, Evan O’Connor, Di Zhang, and Yue Xu. Special thanks to
Sebastian Araya who helped with many of the figures in this book. I wish to thank
my colleagues who have worked with me side-by-side on these projects includ-
ing Aaron Wemhoff, Gang Feng, and Randy Weinstein. Financial support for
our phase change research has come from the Office of Naval Research and the
National Science Foundation. I also wish to thank Prof. Frank Kulacki for pro-
viding the opportunity to write this book. Finally, I wish to thank Paul and Katie
Fleischer for their constant encouragement and support of my efforts.

vii
Contents

1 An Introduction to Phase Change Materials . . . . . . . . . . . . . . . . . . . . . 1


1.1 Phase Change Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Thermal Management versus Thermal Energy Storage. . . . . . . . . . . 3
1.3 Advantages of PCMs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Outline. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Energy Storage Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7


2.1 An Introduction to Energy Storage Applications. . . . . . . . . . . . . . . . 7
2.2 Thermal Management of Electronics. . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Energy Storage in Building Materials . . . . . . . . . . . . . . . . . . . . . . . . 12
2.4 Solar Energy Systems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4.1 Concentrating Solar Power Plants. . . . . . . . . . . . . . . . . . . . . 15
2.4.2 Domestic Solar Applications . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.5 Packed Bed Designs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.6 Heat Exchanger Designs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.7 Textile Designs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.8 Space Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

3 Types of PCMs and Their Selection. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37


3.1 Important PCM Material Properties. . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.2 Organic PCMs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.3 Inorganic PCMs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.4 Metal and Metal Alloy PCMs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

4 PCM Design Issues. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49


4.1 Overview of PCM Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . 49

ix
x Contents

4.2 Enhancement of Thermal Conductivity. . . . . . . . . . . . . . . . . . . . . . . 50


4.2.1 Metallic Inclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.2.2 Macroscale Carbon Inclusions. . . . . . . . . . . . . . . . . . . . . . . . 55
4.2.3 Nanoscale Carbon Inclusions. . . . . . . . . . . . . . . . . . . . . . . . . 58
4.3 PCM Containment Issues. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.3.1 Basic Designs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.3.2 Macro-encapsulated PCMs. . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.3.3 Micro-encapsulated PCMs. . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.3.4 Form Stable PCMs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

5 Fundamental Thermal Analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75


5.1 Introduction to the Analysis of Solid–Liquid Phase Change. . . . . . . 75
5.2 Stefan Problem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
5.3 Advanced PCM Analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

6 Future Directions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
6.1 Development Needs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
6.2 Conductivity Enhancement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
6.3 Specific Heat Enhancement. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
6.4 Latent Heat Enhancement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

Index. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
Chapter 1
An Introduction to Phase Change Materials

1.1 Phase Change Basics

Phase change materials (PCMs) are materials that undergo the solid-liquid phase
transformation, more commonly known as the melting-solidification cycle, at
a temperature within the operating range of a selected thermal application. As a
material changes phase from a solid to a liquid, it absorbs energy from its sur-
roundings while remaining at a constant or nearly constant temperature. The
energy that is absorbed by the material acts to increase the energy of the constit-
uent atoms or molecules, increasing their vibrational state. At the melt tempera-
ture the atomic bonds loosen and the materials transitions from a solid to a liquid.
Solidification is the reverse of this process, during which the material transfers
energy to its surroundings and the molecules lose energy and order themselves
into their solid phase. This can be seen in Fig. 1.1.
The energy that is either absorbed or released during the melting-solidification
cycle is known as the latent heat of fusion. Latent heat is unique in that it is heat
that is absorbed into a material without the material itself increasing in tempera-
ture. It is easy to picture this process by considering the melting of an ice cube.
You can heat the ice cube by exposing it to ambient room temperature conditions,
by heating it with a hair dryer, or by blasting it with a blow torch, but no matter
how much heat flux is supplied to it, that ice cube will not increase in temperature
until the melting process is complete. The latent heat absorbed during the melting
process is referred to as the latent heat of fusion, in order to distinguish it from the
other form of latent heat, the latent heat of vaporization, which characterizes the
change in phase from a liquid to a gas.
In contrast to latent heat, which does not increase the temperature of a m ­ aterial,
sensible heat is that heat which does result in a change in temperature within the
material. A standard continuous heating process may begin with a subcooled solid,
which is heated to the melting point through sensible heating. As the heating

© The Author(s) 2015 1


A.S. Fleischer, Thermal Energy Storage Using Phase Change Materials,
SpringerBriefs in Thermal Engineering and Applied Science,
DOI 10.1007/978-3-319-20922-7_1
2 1  An Introduction to Phase Change Materials

Fig. 1.1  The melting/
solidification process

Fig. 1.2  Standard heating
curve

process continues the solid transitions to a liquid through the latent of heat fusion,
and sensible heat then increases its temperature to the boiling point. Once the boil-
ing point is reached, the liquid transitions to a vapor through the latent heat of
vaporization until the phase change process is complete. Any additional heating
is now in the form of sensible heat which acts to superheat the vapor. This process
can be seen in Fig. 1.2.
It can be seen in Fig. 1.2 that the latent heat of vaporization is a higher energy
process than the latent heat of fusion. Therefore, you might wonder why PCMS
are utilized for their latent heat of fusion rather than their latent heat of vapori-
zation. It is true that in general that the boiling/condensation process absorbs/and
releases more energy, but the density change from a liquid to a vapor is large, and
working with boilers and condensers often requires a significant amount of sup-
port equipment which is not always convenient. There are of course many applica-
tions for boiling heat transfer, but here we will concentrate on the applications for
which a solid-liquid phase change process is most advantageous.
The amount of energy absorption or release during the melting-solidification
cycle is governed by the value of that material’s latent heat of fusion. The latent
heat of fusion is commonly expressed in units of J/g or kJ/kg. Thus the process
is a mass-based process. The amount of energy absorbed by the material during
1.1  Phase Change Basics 3

melting depends solely on the mass of material present in the design. The rate at
which the material melts is governed by the operating conditions. For instance,
water (ice) has a latent heat of fusion of 333 kJ/kg, so 1 kg of ice will require a
heat input of 333 kJ to melt. The rate at which the ice melts depends on the heat
transfer process. How quickly can the heat be transferred into the ice? The applied
heat flux and the temperature difference from the heat source to the melt point will
govern this process. Thus it becomes clear that 1 kg of ice will not increase in tem-
perature if it is melted using a blow torch, rather than being left out on the kitchen
counter, and it will not take more energy to melt, but the ice melted with the
blow torch will certainly melt faster due to the increased heat flux and increased
temperature differential.

1.2 Thermal Management versus Thermal Energy Storage

Energy storage through solid-liquid phase change is inherently a transient process.


The material is either absorbing or releasing energy as its melts or solidifies. Thus
this type of system is not particularly well suited for applications that operate pri-
marily in steady-state conditions. Instead it is best suited for systems that experi-
ence repeated transients, such as on-off cycles, periodic peaking cycles or which
require thermal energy storage for later use.
Systems that operate in off/off or peaking load cycles will use PCMs for what
is referred to as thermal management. In these applications, the energy that is
being released is absorbed into the PCM in order to prevent overheat elsewhere
in the system. When the duty cycle has finished, the stored energy is released to
the ambient and the PCM solidifies in preparation for the next peak energy event.
A common example of this type of system is the on/off or high use/low use cycle
of many portable electronics. As the system transition into high use mode, addi-
tional energy is dissipated when compared to the baseline case. This energy can
be absorbed in a PCM at constant temperature. As long as the PCM melt cycle
is longer than the duty cycle of the electronics, the system should not increase
in temperature. Once the duty cycle ends, the hot PCM dissipates its heat into
the environment as it solidifies. This type of application requires PCMs which
respond well to rapid transients. The heat must penetrate quickly and effectively
into the PCM mass, leading to congruent melting throughout the system. The sys-
tem should also be designed for maximum melting times of the PCM in order to
extend the operational range of the duty cycle as long as possible. This type of
application can also include solid state electronics, battery pack thermal manage-
ment and human comfort systems such as military uniforms or sports equipment.
Systems that require storage of thermal energy for later system usage will use
PCMs for what is referred to as thermal energy storage or TES. In these applica-
tions, during times of excess energy production, some of that thermal energy can
be diverted into a storage system where it is kept for utilization at a later time
period. The storage system features a solid PCM which slowly transitions to a
4 1  An Introduction to Phase Change Materials

Fig. 1.3  Thermal energy storage for solar power plants

liquid at constant temp as the thermal energy flows into the material. The energy is
thus stored in within the material itself, and can easily be extracted by solidifying
the material.
This application is quite different from thermal management applications
because here the thermal energy is the desired product. A common example of this
type of system is a solar energy power plant. Many large scale solar energy plants
provide a surplus of energy at certain times of day. If the plants can be designed
to store the surplus energy for use during times of decreased supply, the plant will
operate more economically and efficiently as seen in Fig. 1.3. Despite the differ-
ence in the operational characteristics when compared to thermal management
applications, TES applications have many of the same PCM requirements. TES
applications do not require the extended melt times of the thermal management
cases, but do require high latent heat materials in order to store as much energy as
possible. As before, it is best if the PCM responds respond well to rapid transients
for better operation during the system switch to stored energy and the heat must
penetrate quickly and effectively into the PCM mass. This type of application can
also include high efficiency heating and air conditioning systems, domestic hot
water systems, and advanced building materials.

1.3 Advantages of PCMs

When compared to the other options available for thermal management and ther-
mal energy storage applications, PCMs are found to have significant benefits. The
main options available for thermal management of transient systems are actively
1.3  Advantages of PCMs 5

cooled air and liquid methods such as fans, cold plates, heat spreaders and heat
sinks. The use of PCMs for transient thermal management has the advantage of
maintaining a constant system temperature throughout the melt process regardless
of applied heat flux. PCMs are lightweight, portable and highly reliable depend-
ing only on the characteristics of the material itself, and do not depend on an
external flow source such as a fan or pump.
The main options available for thermal energy storage include sensible heat
storage and thermochemical storage. Latent heat storage has a much higher energy
density than sensible heat storage, resulting in less required material mass and/or
smaller storage tank volumes. Latent heat storage systems are also easier to work
with than thermochemical storage. The solid-liquid transition results in only a
small density change, resulting in smaller system size and less support equipment
than when attempting to store thermal energy for long term use through the liquid-
vapor phase change process.
PCMs, however, are far from perfect solutions. The detriment most commonly
cited to their greater utilization is that many PCMs do not have high thermal con-
ductivities or diffusivities, preventing rapid system transients. These issues will be
explored in greater depth through this book.

1.4 Outline

This book is designed to provide a comprehensive, although not exhaustive, intro-


duction to phase change materials, their fundamentals and their uses in applica-
tion. Chapter 2 provides an in-depth look at many of the common applications of
PCMs including their use in electronic systems, concentrating solar power plants,
domestic solar thermal energy systems, transient heat exchangers, HVAC sys-
tems, advanced building materials, textiles and space systems. Chapter 3 looks
in depth at the classifications of PCMs, their important thermal properties, and
provides reference tables of common materials. Chapter 4 looks at design issues
and development to improve the thermal characteristics of PCMs through the use
of embedded high conductivity materials. Chapter 5 presents the fundamentals
of the solid-liquid phase change process including the details of the governing
equations and their solutions with different boundary conditions. The book con-
cludes with a look forward to new and necessary developments in the PCM field
in Chap. 6.
Chapter 2
Energy Storage Applications

2.1 An Introduction to Energy Storage Applications

As discussed in Chap. 1, energy storage through solid-liquid phase change is


inherently a transient process and is best suited for systems that experience
repeated transients, such as on-off or periodic peaking cycles, or for those systems
which require thermal energy storage for later use. PCMs are commonly used in
applications for both thermal management and for thermal energy storage.
Interest in PCMs for thermal management of systems can be traced back at least
through the 1970s. NASA in particular was interested in the use of PCMs as what
were then referred to as “thermal capacitors” and PCMs were implemented in several
moon vehicles and in Skylab [1]. The 1977 NASA tech brief “A Design Handbook
for Phase Change Thermal Control and Energy Storage Devices” [2] was one of the
first comprehensive PCM references, and is still widely cited and used today.
During the 1970s and 1980s, interest was also building in the application of
PCMs in solar systems [3–5] for thermal energy storage in both large solar plants,
and in smaller domestic applications such as domestic hot water systems. The
concept of embedding PCMs in various types of building materials, such as wall-
board and floorboards, in order to create houses and offices with lower heating
and cooling loads for greater energy efficiency, also began in the 1970s/80s [6, 7].
Simultaneously, a significant amount of fundamental research was being completed
on PCMs, considering in-depth the melting and solidification processes, and the
roles of conduction and natural convection on the phase change processes [8–11].
With the growth of computing power through the 1980s and 1990s, integrated
circuits began dissipating significant amounts of heat and PCM applications in
the thermal management of high performance, military and consumer electron-
ics came on the scene in the late 1990s [12–14]. More recently, PCMs have seen
application in textile design for energy absorbing clothing for military and con-
sumer products [15].

© The Author(s) 2015 7


A.S. Fleischer, Thermal Energy Storage Using Phase Change Materials,
SpringerBriefs in Thermal Engineering and Applied Science,
DOI 10.1007/978-3-319-20922-7_2
8 2  Energy Storage Applications

This chapter takes a look at some of these more popular applications. The use
of PCM in each application is explained considering the end goal of the design and
its temperature range. With approximately 45 years of PCM usage to consider, this
chapter certainly is not meant to be an exhaustive review of PCM applications, but
instead is meant to illustrate how and why PCMs are being implemented in each
case, with the benefits of PCM implementation and also any design concerns noted
in each case. When available, comprehensive reviews on each topic are cited.

2.2 Thermal Management of Electronics

The design of electronics over the past five decades has closely followed “Moore’s
law” in which processing power doubles approximately every 2 years. This expo-
nential increase in processing power has been a great boon to the field of elec-
tronics, but a great challenge for thermal engineers. Particularly when combined
with a significant decrease in packaging size, the heat transfer aspects of electronic
packaging grow more challenging every year. For reliability reasons, most chip
packages are constrained to operate below 85 °C and all the generated heat must
be dissipated into the environment during both steady-state and transient operating
conditions.
For standard computing systems such as laptops and desktops, the heat loads
can usually be dissipated using a heat sink coupled with a fan, assuming enough
space exists in the casing for the heat sink geometry. High performance comput-
ing systems with higher heat flux loads are increasingly turning to liquid based
cooling systems such as cold plates, which then necessitates the use of auxiliary
support equipment such as pumps, piping and external heat exchangers. But for
portable electronics, one of the largest segments of the consumer electronics mar-
ket, and which includes tablets and smart phones, possible thermal management
solutions are severely constrained by their form factor. The demand for ultra-thin
systems precludes the use of large air-cooled heat sinks or pumped liquid loops.
Fortuitously, most portable electronics are used in on/off or peaking duty
cycles, which makes the use of PCMs for thermal management feasible. Many
tablets and smart phones are in low-power standby mode for most of the day, with
random bursts of activity that cause processing power to peak. For these applica-
tions, PCMs can be used to absorb these bursts of energy and then to dissipate
the stored heat when the peaking cycle has ended. The idea is to have the heat
effectively penetrate the PCM when the peaking cycle begins, melting the PCM
and maintaining a constant operating temperature. The length of the PCM melt
cycle should be matched to common usage time intervals (perhaps 10–30 min).
Once melted, the PCM must shed its heat to the environment as it solidifies and
“recharges” for the next cycle. The use of PCMs in this application is to delay the
onset of steady-state conditions for as long as possible. Once the PCM is fully
melted, if the electronics still remain on, the temperature will rise through sensible
heating to a steady-state operating condition (See Fig. 2.1).
2.2  Thermal Management of Electronics 9

Fig. 2.1  Delay to steady-state of PCM in electronics thermal management

Fig. 2.2  Electronics casing
temperature during peaking
operation with and without
PCM

The use of PCM in this way maintains a more constant temperature of the
electronics in peaking operation (see Fig. 2.2), and is a passive thermal solution
with no mechanical working parts like fans or pumps, thus increasing reliability.
In this case the PCM is used in a thermal management application, not an energy
storage solution, since the stored heat is not used productively elsewhere in the
system. The PCMs used in these applications typically have melt temperatures
between 36 and 56 °C in order to keep the junction temperature well below the
85 °C allowable for integrated circuits. For portable electronics, it is important
not only to keep the junction temperature low, but also the casing temperature
low in order to protect the user from burns. In general, casing temperature must
be maintained below 40–45 °C for safe usage. PCMs based on paraffin are most
commonly used in these systems although liquid metal blends are sometimes
implemented.
10 2  Energy Storage Applications

Electronics thermal management PCM research has been ongoing since the
1990s and these types of solutions are of great interest to today’s electronics
manufacturers. Many companies are actively designing PCM based thermal solu-
tions and a number of small companies have PCM based heat sinks and spreaders
already on the market. By taking a brief look at some of the literature in the field,
the usefulness of this technology can be seen.
The initial work in this field focused on proving that PCMs could effectively
act to suppress temperature spikes and maintain consistent junction temperature in
operational ranges similar to those occurring in electronic systems. For instance,
in the late 1990s Pal and Joshi [13] completed a numerical study of the use of
PCM for passive thermal control in electronics systems during variable power
operation. They used two different PCMs, eicosene (a paraffin) and a eutectic
alloy of Bi/Pb/Sn/In. It was shown that both PCMs effectively controlled the sys-
tem temperature. Vesligaj and Amon [14] also looked at time varying workloads
on electronic systems, both experimentally and numerically. The duty cycle in this
case was an initial ramp up of 45 min at 10 W of power, followed by 30 min off
and 15 min on in repetitive cycles. It was seen that the use of PCM damped out the
system temperature swings resulting from the power cycles, maintaining a casing
temperature of 31 ± 1 °C compared to 37 ± 7 °C without any PCM, for a signifi-
cant overall temperature and temperature swing reduction.
The effectiveness of PCM thermal management as compared to copper heat
spreaders was illustrated by Krishnan et al. [16] who performed a theoretical anal-
ysis which compared the thermal performance of different PCMs to that of a con-
ventional copper heat sink. The analysis considered the effects of conduction and
phase change through the different materials when subjected to large heat loads of
between 300 and 600 W. The PCMs considered included triacontane (melt temper-
ature  = 65 °C), aluminum foam impregnated with triacontane, and two metallic
PCMs—a Bi/Pb/Sn/In alloy and a Bi/In/Sn alloy which have melt temperatures of
57 and 60 °C respectively. It was found that that the PCMs were significantly bet-
ter at controlling the junction temperature (10–20 °C cooler) than the traditional
copper heat sink and consistently extended the time to achieve steady state. This
work noted the need to enhance the thermal conductivity of most organic PCMs
for effective operation, which was done in this case with aluminum foam. This
option will be discussed in depth in Chaps. 3 and 4.
These works, and others, fully established the potential impact of PCMS in the
thermal management of electronics. As such, many researchers have studied the
direct implementation of PCM in electronics applications. The feasibility of using
a PCM for transient thermal management of a cellular phone was investigated by
Hodes et al. [17]. In this project, the transient response of a mock handheld phone
was determined using two types of PCM: tricosane, a common paraffin wax with
a melt temperature of 48 °C, and Thermasorb-122, a commercially available PCM
encapsulated in a pellet-like plastic shell with a melt temperature of 50 °C. It was
determined that for low power loads even small masses of PCM can substantially
increase the operational time for a handheld phone before overheat occurs. For a
heat load of 3 W, the presence of the PCM doubled the overall usage time of the
2.2  Thermal Management of Electronics 11

device before the casing reached the peak operational temperature. In most port-
able electronics, the casing temperature limit is reached long before the junction
temperature limit.
A similar study [18] looked at a slightly larger system (97 mm by 72 mm by
21 mm) which is certainly too thick to mimic today’s smart phones, but does pro-
vide some insight nonetheless. Four different plate fin heat sinks were designed
for the system. Three of these featured a varying numbers of fins with n-eicosane
with a melt temperature of 36.5 °C filling the space between the fins. The fourth
was a standard air cooled heat sink. The heat sinks were subjected to different duty
cycles of 3–5 W as follows: Light usage (On for 5 min, Off for 50 min, On for
5 min), Moderate usage (On for 15 min, Off for 30 min, On for 15 min) and heavy
usage (On for 25 min, Off for 10 min, On for 25 min, Off for 10 min). Without
the PCM, the casing temperature quickly exceeded its temperature limit of 45 °C
while the PCM filled heat sinks featured greater operational times and lower
device temperatures. The heat sinks the greatest number of fins exhibited the low-
est device temperatures. This was because the fins provided a direct path for heat
penetration into the PCM mass pointing to the benefits of enhancing the thermal
conductivity of paraffin PCMs.
The results of this test also showed that it takes significantly longer to solidify
the PCM than to melt it, indicating the recharge time is a limiting factor in many
applications. This is due to the conduction-dominated nature of the solidification
process as compared to the natural convection-dominated nature of the melt pro-
cess. As the PCM melts, it exhibits density differences which induce natural con-
vection enhancing the melt process and speeding its progression. The low thermal
conductivity of the PCM and the slower nature of the conduction-dominated freez-
ing process can potentially lead to long recharge times which must be considered
during the design process.
For today’s thin form factor electronics, the use of PCMs embedded in heat
spreaders may be an effective solution. The use of very thin layers of PCM mini-
mizes the solidification time while also meeting geometric constraints. For exam-
ple, a PCM based heat spreader was fabricated from electro-deposition of metal
over a template of spherical microcapsules (<100 μm in diameter) containing
PCMs [19]. The authors refer to this as a microporous metal matrix (MMM)-PCM
composites. The paraffin used in this application was SunTech P116 with a melt
temperature of 47 °C and the deposited metals were copper and aluminum. The
resulting heat spreader can be very thin depending on how many layers of PCM
microcapsules are used. A COMSOL simulation of this design indicated that that
it was effective at maintaining the device casing temperature under the temperature
limits and that the temperature of the MMM-PCM composite heat spreader after
25 min was 30 °C below the temperature of a comparable copper heat spreader.
Of course portable electronics are not the only type of electronics that work
within peaking or on/off duty cycles. In fact many power electronics systems also
work this way. Electronics power supplies are also beginning to draw attention
as potential PCM thermal management applications, particularly energy–dense
Li-ion battery packs. Li-ion packs, such as those used in hybrid vehicles generate
12 2  Energy Storage Applications

heat during their discharge cycle, yet the lifespan of the battery pack is highly
dependent on maintaining a consistent operating temperature. In one study [20],
composite blends of paraffin and expanded graphite paraffin with different phase
change temperatures (36, 44 and 52 °C), packing density and mass percentage of
paraffin were analyzed for the effective thermal control of four simulated battery
packs. It was found that the PCM composites effectively maintained a consistent
battery pack operating temperature avoiding overheat.
From this overview it can be seen that PCMs have been shown to be an effec-
tive solution for the transient thermal management of different types of electronic
systems. The primary advantages to using PCMs in these applications are increased
reliability due to the passive nature of the melt/solidification process and the suit-
ability for low form factor systems. The main challenge to their greater implemen-
tation is the low thermal conductivity that many PCMs with melt temperatures in
this range exhibit, although as shown, there are ways to mitigate the impact.

2.3 Energy Storage in Building Materials

Energy storage has long been a part of the selection of building materials. From
the earliest days, man has wanted to design comfortable dwellings that absorb heat
during the day, preventing overheating, and which retain as much heat as possible
at night, maintaining a comfortable temperature within. The use of material such
as adobe is a prime example. These materials store significant amounts of sensible
heat during the day, and release it slowly during the cooler night hours creating a
comfortable abode in desert environments. Adobe housing has been popular for
millennia, and among the most famous adobe dwellings is Cliff Palace, in Mesa
Verde National Park in Colorado (see Fig. 2.3). Adobe houses are still a popular
option in the American Southwest today.

Fig. 2.3  Mesa Verde’s Cliff Palace sandstone and adobe dwelling


2.3  Energy Storage in Building Materials 13

The interest in PCMs in building materials is in the ability to harness energy


storage through the latent heat phase transition instead of through sensible heat-
ing. The primary benefit to this in building materials is the higher energy density
available when storing energy through phase transition, meaning that more energy
can be stored in a constant volume. By storing energy through the phase transi-
tion, the overall HVAC load of the structure can be reduced. This is related to the
constant temperature of the phase transition. When solar energy is incident on sen-
sible energy storage materials, the temperature of the materials rises. For buildings
this means that in warm weather months the HVAC system is needed to maintain
comfortable temperatures within. However, with PCM-based building materials,
the temperature will only rise until the point of phase transition is reached. At that
point, continued incident energy is absorbed through the latent heat of fusion, and
the temperature remains constant. If this can be matched to the building comfort
set point, HVAC loading can be significantly reduced. This works particularly
well as noted in the high desert environments where it can be hot with high solar
insolation during the day and yet also have cool nights during which the heat is
released from the PCM.
Not surprisingly then, the main areas of emphasis for PCM implementation for
energy storage in building systems are in PCM-based gypsum board. In most of
these systems, the PCM transitions from solid to liquid in the range of 20–30 °C.
The method of containment of the PCM in the gypsum board varies from design
to design, and a few of these applications are highlighted here. Currently there are
several commercial PCM gypsum boards available, many of which use an encap-
sulated PCM supplied by BASF which melts at 23 °C.
There are many research teams which have studied the use of PCMs in wall-
board to quantify its effectiveness. For instance, in an early paper by Feldman
et al. [21], PCM based on methyl palmitate with a melt range of 23–26 °C and a
latent heat of 180 kJ/kg was impregnated at 25 wt% into a standard gypsum wall-
board by direct immersion of the wallboard in the liquid PCM. For this material,
the surface tension was such the liquid PCM wicked into the pores in the wall-
board, and upon repeated melt/freeze cycles, the molten liquid PCM was effec-
tively constrained via capillary forces within the wallboard without any seepage
of PCM. Thermal property testing of the composite material showed that over a
3.5 °C interval the total storage capacity was about twelve times higher than the
storage capacity of wallboard alone, showing a significant increase in energy
density.
Full-scale room testing was done by Athienitis et al. [22]. In this study, a gyp-
sum board with 25 wt% PCM was again made using direct immersion of wall
board in liquid PCM, in this case butyl stearate with a phase change transition
between 16–21 °C. The PCM wallboard was installed in an existing room test bed
2.82 m wide by 2.22 m long and 2.24 m tall. The experiments were run under
simulated winter conditions with the outdoor ambient set as cold as −25 °C, and
the indoor air temperature set at 23 °C from 6:00 a.m. to 5:00 and to 16 °C outside
of those hours. The room had a single window (1.08 m2) which allowed for inso-
lation. Previous testing with this same test bed under sunny day conditions had
14 2  Energy Storage Applications

shown that the room temperature could reach 30 °C, while under the exact same
conditions the PCM-based wall board passively controlled the temperature of the
room to 26 °C. During the discharge cycle at night, the surface temperature of the
PCM gypsum board was 1–1.5 °C higher than the standard wallboard, clearing
showing direct energy savings both during the day and at night.
Some more recent studies, such as Zhang et al. [23] have used PCM micro-
encapsulation instead of direct immersion in liquid PCM. In microencapsulation,
the PCM is contained within a microscale polymeric bead. This encapsulation
removes the dependence on capillary action to contain the liquid PCM within
the wallboard. Instead the PCM stays safely with the shell at all times, prevent-
ing leakage and also preventing any off-gassing during melting. In this study
[23], the PCM capsules were combined with gypsum powder and glass fiber to
form gypsum boards by compression molding. The PCM was octadecane with a
melt range of 26–28 °C. Sample composites with ratios of PCM/Gypsum from
30/70 up to 60/40 were fabricated and tested. A single 50/50 composite gypsum
board was found to absorb energy and moderate temperature rise for periods of
up to 48 min, while also reaching lower final temperatures than standard gypsum
board and eliminating entirely any possibility of PCM seepage. A more detailed
review of the use of microencapsulated PCMs in building materials is provided by
Tyagi et al. [24] and a review of PCMs of various types in building materials by
Parameshwaran [25].
Of course wallboard is not the only building material into which PCMs can be
dispersed for energy storage. Concrete is another such example. Han et al. [26]
studied microencapsulated PCMs dispersed within cement. The microencapsula-
tion is used in this case because liquid paraffin is extremely difficult to disperse in
cement which can lead to non-uniformity throughout the composite. The microen-
capsulated PCM beads are much easier to disperse within the cement. The PCM
used in this case is a commercially available Micronal DX technology made by
BASF with a melt point of 26 °C. In this case, carbon nanotubes are also added
to the cement to offset the reduction in strength and thermal conductivity that can
result from the PCM addition. The composites featured 5 % PCM and 1 % CNT
by weight. The composite was used to fabricate a scale-down house-sized building
model, with a control model built with plain mortar as a reference. The building
model fabricated from the PCM cement showed a reduction in internal tempera-
ture of almost 7 °C under outdoor test conditions.
The use of PCM in roofing materials has also been investigated. Kosny et al.
[27] studied an experimental roof design which incorporated a PCM energy stor-
age system for use with roof mounted photovoltaics (PV). The roof design was
a composite made from layers (in order) of roof decking, macroencapulsated
PCM, fiberglass insulation and metal panels with pre-installed PVs. The mac-
roencapsulated PCM consisted of an organic PCM with a melt temperature range
of 26–30 °C in plastic film pouches with dimensions of 4.4 × 4.4 by 1.3 cm.
Traditional roofing materials (standard shingle roof) served as the experimen-
tal control. Data on the performance of the roofing systems were collected con-
tinuously over a 12 month period. Overall, the PCM roofing system was shown to
2.3  Energy Storage in Building Materials 15

offer benefits on about 2/3 of the days of the year. The effectiveness was reduced
because during some winter periods the PCM remained entirely solid with no
melting, and during certain summer periods the PCM remained entirely liquid
with no solidification. The remainder of the time it cycled through phase change
for at least some of the day. During the winter season as a whole, the PV-PCM
roofing system reduced heating load by 30 %, and during the summer the average
cooling load was reduced by 55 %.
It can be seen from this brief overview that PCMs have been shown to be an
effective way to reduce energy costs when embedded in building materials. The
PCMs can be embedded by direct immersion, macroencapsulation or microencap-
sulation, depending on the host material and application. The primary advantages
to using PCM in building materials is their ability to store energy without tem-
perature increase, thus mitigating temperature rise and thus HVAC loading within
the building. Several commercial products are now available, and the only barrier
to their greater implementation is currently cost and wide scale availability. These
materials can be further refined any optimized by addressing issue of melt temper-
ature range to avoid situations where the PCM remains either fully liquid of fully
frozen for days at a time, as well as issues of weight and for concrete, structural
integrity.

2.4 Solar Energy Systems

2.4.1 Concentrating Solar Power Plants

An obvious drawback of solar power systems of any type is the limitation of the
effectiveness of the technology to periods of high radiant solar energy. This is par-
ticularly an issue for large commercial solar power plants, as electricity demand
is not limited to daylight hours. The use of phase change materials for thermal
energy storage (TES) in these applications can extend the usefulness of the tech-
nology so that benefits can be provided even where there is low or no direct insola-
tion. In fact, for large solar power plants, fully integrated TES is necessary for the
design of economically viable plants in order to reduce reliance on supplemental
fossil fuel burners.
Commercial solar power plants are designed using the concept of
Concentrating Solar Power (CSP). In these plants, sunlight is reflected and con-
centrated using mirrors and then used to heat a carrier fluid. There are four main
types of concentrator designs: parabolic trough, linear Fresnel, power tower (helio-
stat) and dish-sterling. A review of these designs is available in Barlev et al. [28].
Parabolic trough is the most mature technology and is being used in a number of
operating power plants around the world including Andasol 1–3 in Spain (three
50 MW power plants) and the 250 MW Solana Generating Station in Arizona. An
example of parabolic trough technology is shown in Fig. 2.4. In this image, the
thermal receiver is supported above the concentrating mirrors. The receiver is a
16 2  Energy Storage Applications

Fig. 2.4  Parabolic trough
concentrating solar collectors
with heat transfer fluid in
the tubing absorbing thermal
energy (http://energy.
gov/eere/renewables/solar)

black pipe encased in a vacuum tube to reduce convective losses. A high tempera-
ture, high pressure heat transfer fluid (HTF) circulates through the receiver pipes.
Depending on the design of the system, the HTF fluid may serve as the heat source
in an evaporator, creating steam which powers a stream turbine which drives a
generator, or the HTF may directly vaporize as it passes through the solar field
and then pass straight through the turbine without an intermediate heat exchanger
(known as Direct Steam Generation—DSG). In either design, during periods of
high insolation, it is possible to absorb more solar thermal energy into the HTF
than is necessary to power the turbine. This “excess” solar thermal energy can be
stored using sensible or latent heat in storage tanks as shown in Fig. 4.2.
Thermal energy storage in CSP plants has been studied by Denholm and Mehos
[29] who looked at the load shifting possible by storing solar thermal energy. They
point out that load shifting is beneficial both in terms of production of energy
during off-peak solar periods, and in the ability to improve system flexibility by
reducing constraints of ramping and minimum generation levels. Ramping and
minimum generation levels are related to the need for conventional fossil-fuel
powered systems to ramp up rapidly to address the decreased solar output during
peak evening. In order to meet this ramp rate, it is common to operate the fos-
sil fuel systems at part-load even during periods of solar high output. However
if stored thermal energy can be dispatched to meet the peak demand in the early
evening instead, the need for auxiliary systems will be minimized [29] (Fig. 2.5).
Sioshani and Denholm [30] also considered the economic impact of TES on
CSP plants, particularly in the southwestern United States. In this case they exam-
ined the load shifting capability of TES systems in the context of shifting gen-
erating capability to peak electricity cost periods, thus maximizing profit. Their
analysis of some Texas locations indicated that due to the matching of peak insola-
tion time with local electricity demand patterns that CSP alone is between 7 and
35 % more valuable than the average price of electricity in the region and that
adding TES increases this value by another 7–16 %. This economic benefit comes
from the load matching allowing for a large solar collection field and from the
higher sale prices from load shifting. From these in-depth studies it is clear that
that additional of thermal energy storage to CSP plants makes economic sense.
2.4  Solar Energy Systems 17

Fig. 2.5  An example of a direct steam generation concentrating solar power plant with thermal
energy storage

In fact, most operational plants, including Andasol and Solana have thermal
energy storage capability. Andasol has 7.5 h of thermal storage capacity while
Solana has 6 h. Both of these plants use what is known as the two-tank molten
salt system. In the two-tank molten system a heat exchanger is located between
the two tanks with the HTF flowing on one side of the exchanger and the storage
medium (molten salt) on the other side. During the energy storage cycle, some of
the HTF from the solar is diverted to this exchanger where it transfers energy to
the molten salt. In this case, the salt flow originates in the “cold” tank and flows
through the heat exchanger where it absorbs solar thermal energy and then into
the “hot” tank where it is stored [31]. During the energy discharge cycle, the HTF
and molten salt flow paths are reversed. The salt gives up its energy to the HTF as
it moves from the hot tank through the heat exchanger into the cold tank, and the
now hot HTF is used in the power cycle. While these systems have seen success,
there is significant cost inherent in using two storage tanks, and the energy density
of these storage systems is low as the salt remains in the liquid phase at all times.
The use of PCMs in these applications can thus reduce tank number (to one), size
and installation costs, creating an economic benefit.
Molten salts are commonly used in these applications because of their high
operating points. These materials have melting points from around 300 °C to over
800 °C. The HTF in parabolic trough and linear Fresnel system can reach around
300–400 °C in the receiver, while heliostats receivers can operate in excess of
2000 °C [28]. Salts are well suited for these operational ranges, but suffer from
a few drawbacks including high corrosiveness and low thermal conductivity. The
primary issue with low thermal conductivity is the need for quick charge and dis-
charge of energy as the HTF flows through the storage medium. In a few cases,
liquid metal alloys may be used instead of molten salts. Adinberg et al. [32] used a
18 2  Energy Storage Applications

70 wt% Zinc/30 wt% Tin alloy at TES temperature of 370 °C. The alloy exhibited
high chemical stability and thermal conductivity of 50 W/m-K in the liquid state,
an increase of about two orders of magnitude over molten salts. Both molten salts
and liquid metals are discussed in more depth in Chap. 3.
An economic analysis of PCM thermal energy storage for CSP applications
was completed by Robak et al. [33]. In this work they compared the costs and per-
formance of the standard two-tank sensible heat storage model to a PCM system.
In this case, the low thermal conductivity of the PCM was offset by using ther-
mosyphons (gravity heat heat pipes) embedded within the PCM to promote heat
penetration. The analytical model developed for the PCM system showed that for
9 h of storage for a 50 MW generating plant, the TES tank must be 17,464 m3
with 30,000 tons of PCM and 3,300,000 thermosyphons. In comparison, the sensi-
ble heating system requires two tanks, each 23,856 m3 with 45,000 tons of molten
salt. Thus the PCM TES system reduces the required overall tank volume by
approximately 65 %, and reduces the TES medium mass by approximately 30 %
[33]. The PCM system has additional costs for the thermosyphons, and the HTF
channels within the tank but the sensible heat system requires a molten salt-HTF
heat exchanger and molten salt pumps. The authors estimate that overall the PCM
TES system will cost 15 % less than the two tank system.
A detailed economic analysis of the benefits of PCM in CSP applications was
also undertaken by Nithyanandam and Pitchumani [34]. In this analytical study
they used the Levelized Cost of Electricity (LCE) metric to compare CSP plants
operating on either Rankine or s-CO2 cycles with integrated TES. Three different
TES methods were considered—the standard two tank sensible heat system, PCM
with embedded heat pipes, and encapsulated PCM in which the tank is filled with
spherical capsules containing PCM. The PCM used in the Rankine cycle system
was 60 % NaNO3/40 % KNO3, known as solar salt while the PCM used with the
s-CO2 power cycle was KCl/MgCl2.
The total cost of the encapsulated PCM storage system was calculated as the
sum of the costs for the HTF, PCM, container, encapsulation, and overhead costs
consisting of 10 % of the storage material, container and encapsulation costs. The
total cost of the heat pipe PCM system was calculated the same, with the heat
pipe cost replacing the encapsulation cost. Using this cost estimation plus known
data on the costs of two-tank systems, the lowest cost was shown to occur with
the encapsulated PCM for both the s-CO2 and the Rankine cycle systems. For the
s-CO2 cycle system, with no TES the LCE was 11.07 ₵/kWh, with sensible heat
storage it was 6.49 ₵/kWh, with the heat pipes based system it was 5.77 ₵/kWh
and it was only 5.37 ₵/kWh for the encapsulated PCM, a reduction of 48 % from
the no-TES case and a reduction of 11 % over the sensible heat storage case.
Similar reductions in costs were seen for the Rankine cycle systems [34].
It can be seen from this brief overview that the storage of thermal energy has
been shown to be necessary in the implementation of large scale concentrating
solar plants. Although the existing CSP plants use sensible heat storage with a
two tank system, it is clearly shown that the use of PCM can lead to significantly
lower capital and operating costs. The typical PCMs used in these applications are
2.4  Solar Energy Systems 19

inorganic salts which melt in the range from 300–800 °C. These PCMs tend to be
corrosive and have low thermal conductivities but it was shown that this can be
offset with the use of embedded heat pipes or thermosyphons. In certain applica-
tions liquid metals may be used instead.

2.4.2 Domestic Solar Applications

While the large CSP plants certainly have significant technical and economic
incentives to implement PCM thermal energy storage systems, smaller scale
solar systems can also reap some benefits from TES. For example, solar ther-
mal systems can be used by small businesses and homes for hot water produc-
tion and for heating systems. A small scale solar hot water system with energy
storage can be seen in Fig. 2.6. These systems feature a flat plate solar collector,
typically mounted on the roof, which features a heat transfer fluid passing through
the receiver tubes. The receiver tubes are isolated within an enclosure with a glass
cover plate. The enclosure may be evacuated to prevent convective losses. In many
ways this is similar to the CSP solar field, but without the concentrators. The lack
of concentrators means that the HTF will not reach the high temperatures charac-
teristic of CSP. As such the fluid can’t be used to create vapor and drive a power
system, but is hot enough to provide the heat source for a domestic hot water tank.
As with CSP, the effectiveness of the system is limited to daylight hours, but the
solar thermal system can be designed to store extra heat using PCM in the stor-
age tanks for the overnight hours, greatly reducing dependence on supplemental

Fig. 2.6  Domestic solar hot water heating system with PCM thermal storage
20 2  Energy Storage Applications

natural gas or electrical heating. For instance, Sole et al. [35] designed a domestic
hot water tank with a coiled heat exchanger filled with PCM in the top third of a
thermally stratified tank. The PCM selected had a melt temperature of 58 °C and
was a PCM–graphite compound of about 90 vol% sodium acetate trihydrate and
10 vol% graphite. The mass of PCM in this case was only 4.9 kg, compared to the
287 kg of water in the tank. The PCM tank stored approximately 3 % more energy
than a comparable water only tank, proportional to the extra energy which could
be stored in the PCM mass. An exergy analysis showed the PCM tank has an exer-
getic efficiency of 95 %, compared to 85 % for the water tank, illustrating that the
energy stored in the PCM tank was of higher useful quality. Several other appli-
cations of PCM thermal energy storage in combination with solar thermal heat-
ing systems for residential and university buildings are presented in a review by
Kenisarin and Mahkamov [36].
Thermal energy storage can also be used with solar thermal systems intended
to heat air (instead of water) in order to eliminate or reduce the size of HVAC
systems. This is particularly useful for solar heating in greenhouse systems. For
example, a 180 m2 greenhouse facility with 27 m2 of flat plate solar air collectors
[37] used 6000 kg of paraffin wax with a melt range of 48–60 °C as the medium
for solar energy storage. A steel tank 1.7 m in diameter and 5.2 m long served as
the latent heat storage unit (LHS). While the system was shown to be effective and
energy efficient, an exergy analysis found a low net exergetic efficiency, highlight-
ing an issue with the quality and usefulness of the stored heat in this case.
PCM thermal energy storage can also be implemented as part of a solar-aided
heat pump system to allow operation during the night and on cloudy days [38].
During times in which solar insolation occurs, the thermal energy is transferred
to and stored in the PCM tank and when space heating demands occur they are
met using the tank’s stored energy. In this application, a parametric analysis was
completed to determine the effect of PCM choice on the solar energy storage time.
The PCMs considered were calcium chloride hexahydrate (CaCl2 · 6H2O) with a
melt temperature of 29 °C, paraffin with a melt temperature of 47 °C and sodium
sulphate decahydrate (Na2SO4  · 10H2O) with a melt point of 39 °C. The lower
melt point materials melted more quickly, resulting in a faster charge of the sys-
tem, but that was offset by a lower quality of the stored heat at the lower tem-
perature. A similar system was experimentally tested for a flat plate solar collector
powered heat pump system which was used to heat a 75 m2 laboratory building.
CaCl2  · 6H2O served as the energy storage medium [39]. At the installed loca-
tion in Trabzon, Turkey, the stored energy met 60 % of the heating load in March,
100 % in April and provided a surplus in May.
Of course, if PCM thermal energy storage units are useful for solar assisted
heat pump heating systems, then they are also going to be effective with ground-
source heat pump heating systems. This was shown in practice using a ground
source heat pump with thermal storage in CaCl2 · 6H2O to provide heating for a
30 m2 glass greenhouse [40]. Testing over a period from October to May, it was
found that the heat pump produced a temperature increase of 5–10 °C, and the
latent heat storage provided an additional 1–3 °C of auxiliary heating.
2.4  Solar Energy Systems 21

This overview of smaller industrial and domestic thermal energy storage appli-
cations shows that although the economic benefits are reduced when compared
to large CSP plants, there is still a role for thermal energy storage in smaller
domestic scale solar thermal heating systems. The lower temperatures for these
solar applications (50–120 °C) necessitate a different choice of PCM. The typi-
cal PCMs used in these applications melted in the range from 29–55 °C and com-
monly included CaCl2 · 6H2O and paraffin wax. These PCMs have been shown to
be effective when used with domestic solar hot water, solar assisted heat pumps
and ground source heat pump systems.

2.5 Packed Bed Designs

Both large scale and domestic scale solar systems are typically arranged such
that the energy storage is implemented using PCMs contained within indus-
trial tanks of various sizes. Although the use of heat pipes to transfer and extract
stored energy from the tanks was discussed in Sect. 2.4.1, in many other cases
these tanks are arranged in packed bed designs. Packed bed designs typically fea-
ture large spherical capsules containing PCM closely packed within the tank. The
heat transfer fluid flows in, around and over the spheres as it passes through the
tank, as seen in Fig. 2.7. In charging flow, a hot fluid will pass through the tank,
exchanging heat with each sphere, which contains PCM in the solid phase. The

Fig. 2.7  Packed bed design


with PCM spherical capsules
22 2  Energy Storage Applications

energy exchange through the capsule shell leads to melting within and energy
storage within the capsule. For energy discharge flow, the direction of flow is
reversed within the tank. Cold fluid now flows through the tank, which warms as it
passes over the hot capsules which contain liquid phase PCM. Heat is exchanged
from the hot capsule to the colder fluid and the PCM transitions to solid phase
as it exchanges energy. Advantages to packed bed systems include the high heat
exchange contact area, and the ability to exchange energy in both charging and
discharge flow in one single tank, significantly reducing costs over the two tank
design. The containment of the PCM within the spherical capsules also leads to
simple and reliable implementation of the system.
Packed beds designs have been around for decades, but in most cases the paced
bed material does not change phase, featuring only sensible heat storage. The solid
material in these designs is a a material with a high specific heat, which can store a
significant amount of energy through increases in temperature. Common sensible
heat packed bed storage materials include concrete, rocks and masonry. If PCMs
are used instead of sensible heat storage material, smaller volumes are required to
store the same amount of energy through latent heat, thus reducing capital equip-
ment costs.
The influence of the specific heat and latent heat on the thermal response of a
packed bed system was investigated by Arkar and Medved [41]. The packed bed
in this study was a 1.53 m tall tank with a diameter of 0.34 m filled with 35 rows
of 50 mm diameter spheres packed in a rhombic arrangement, creating a system
with an average porosity of 0.388. Each sphere had a wall thickness of 1 mm and
was filled with a paraffin based PCM. Both experimental and numerical models of
this system were completed. It was found that the rate at which energy charges or
discharges from the system is primarily dependent on the specific and latent heat
of the selected PCM, and that the rate of flow of heat transfer fluid through the
packed bed has a smaller influence on the thermal performance. Thus the PCM
utilized must be chosen carefully to ensure the best results for a given application.
The effect of the convection rate on the outside of the capsule was quantified
by Lee et al. [42] who used several different numerical models validated against
experiments. The three models all used a unit cell approach to the packed bed
design as seen in Fig. 2.8 The boundary conditions were applied differently in
each case, with the first case using a constant temperature boundary at the edge
of the PCM, the second a constant temperature boundary at the outer capsule edge
to include the influence of the shell material on the thermal performance while the
third used a convection boundary condition on the shell outer edge, which is most
similar to the actual experimental conditions. The results showed that the inclu-
sion of both convection and conduction through the capsule wall slowed down the
thermal response of the system. The melting process was 14 % slower for this case
when compared to the isothermal boundary condition at the PCM edge. The third
model matched the experiments most closely, and showed that as the melting pro-
cess is slowed, the thermal response of the system is affected by the PCM material
choice, but also to a lesser degree by the capsule material and convection rate.
2.5 Packed Bed Designs 23

Fig. 2.8  Packed bed numerical analysis unit cell

As the PCM material properties were found to have the greatest effect on
packed bed thermal performance, several other researchers focused on the effect
of various thermal property enhancements on packed bed designs, including the
effect of dispersing high conductivity particles within the PCM capsule [43]. This
study developed a numerical model that assessed the impact of macroscale high
thermal conductivity inclusions within the PCM on the energy charge and dis-
charge rates. The thermal conductivity changes were modeled using an effective
thermal conductivity model for the composite material which was a function of the
particle conductivity, PCM conductivity and the particle volume fraction.
The packed bed was modeled in discharge flow during which the PCM slowly
solidified as dissipated heat into the fluid flowing through the bed. The high con-
ductivity particles were not found to exert much influence on the initial stage of
discharge during which the superheated liquid PCM approached the solidification
point. Once solidification began however, and the energy transfer became conduc-
tion dominated, the particles exerted a strong influence on energy discharge. The
addition of the particles to PCM resulted in quicker energy discharge, leading to a
faster warming of the heat transfer fluid and a faster solidification rate of the PCM.
This increased energy transfer rate continued throughout the solidification process
and also through the supercooling of the solid PCM. As the supercooling process
is completely conduction based, increases in thermal conductivity are particularly
significant.
The particle fraction was found to affect the rate of energy discharge with
higher particle fractions leading to faster solidification of the PCM. Particle frac-
tions of 10, 20 and 30 % decreased the solidification time by 8.7, 17.5 and 26.2 %
respectively.
24 2  Energy Storage Applications

The effect of particle conductivity was less significant than the particle frac-
tion. Increases in conductivity ratio of particle to PCM (kparticle/kPCM) from 1 to 10
were found to be significant, but further increases from 10 to 50 had much lower
effects on discharge rate. Thus the type of particle chosen is less significant than
the loading fraction.
This overview of packed beds with thermal energy storage shows that the use
of packed beds offers significant benefits in the application of PCMs for energy
storage. The systems are simple to install and offer a lot of surface area which can
lead to enhanced heat transfer to the operating fluid. Ongoing work on optimiz-
ing the performance of packed beds shows that the choice of PCM has the strong-
est influence on the charging and discharging rates. While PCMs for packed beds
should be selected primarily for appropriate melt temperature range, improve-
ments in thermal properties using high conductivity inclusions can improve ther-
mal performance. The rate of flow through the packed bed was found to exert a
lower influence on overall system performance.

2.6 Heat Exchanger Designs

The previous section illustrated the possible benefits of PCM thermal energy stor-
age in commercial and domestic scale solar systems. In many of these systems, the
PCM is implemented in a heat exchanger type design where the heat from the solar
field is used to charge the storage tank. In fact, PCM energy storage heat exchang-
ers can be useful in any system in which an excess heat capacity can be stored for
future use through heat exchange from a process fluid into a PCM storage system.
These designs often take the form of a shell and tube heat exchanger system
with the PCM on the shell side as seen in Fig. 2.9, although some designs have
the PCM on the tube side. The choice of PCM material in these applications will
depend on the operating temperature range of the system and can include molten
salts for high temperature waste heat recovery and paraffin for lower temperature
domestic heating systems.
Here we will present some of the design and application issues involved with
using PCMs in heat exchangers. There have been numerous studies of PCM in
heat exchanger designs, and there are a couple of comprehensive reviews avail-
able. Jegadheeswaran and Pohekar [44] review a number of shell and tube designs
within the context of a larger review of PCM melting, solidification and heat

Fig. 2.9  PCM on the shell


side of a shell and tube heat
exchanger
2.6  Heat Exchanger Designs 25

transfer enhancement methods and Al-Abibi et al. [45] include a section on PCM
usage in heat exchangers within the context of a review of thermal energy storage
in HVAC systems.
The effective application of PCM in a shell and tube exchanger for thermal
energy storage requires a low thermal resistance between the heat transfer fluid in
the tubes and the PCM in the shell. When operating in charging flow, with hot HTF
acting to melt the shell side PCM, it is desirable to transfer as much heat as possible
as the fluid flows through the tubes. In this case, with melting PCM, the heat trans-
fer will be convection dominant on the outside of the tubes, as natural convection is
induced in the shell side. When operating in discharge flow, with the heat extracted
from the molten PCM, the heat transfer will be conduction dominant as the PCM
solidifies around each tube. Much of the design work in this application has con-
sidered methods to reduce thermal resistance in both melting and solidification in
order to most effectively transfer heat in and out of the PCM to the flowing HTF.
Various different shell and tube side modifications have been analyzed for their
effects on thermal resistance. For instance, the use of longitudinally internally
finned tubes has been considered as a method to induce tube-side turbulence and
create a lower total thermal resistance to the shell [45]. The design was analyzed
numerically during the melting/charge phase. The addition of internal tube fins was
found to significantly increase the rate of PCM melting, and thus decrease charging
time. The melting volume fraction, which reflects the progression of the melt at a
given point in time, increased with the fin thickness, height and number of fins.
The use of high conductivity particles dispersed within the PCM has also been
found to reduce thermal resistance during melting of a shell-side PCM [46]. An
analytical method was developed to study the melting progression in the PCM
and to determine the optimum fraction of particles for maximum energy storage.
The addition of the particles in the PCM was found to reduce the thermal resist-
ance and promote a faster charging (melting) time. However, the inclusion of the
particles displaced some PCM mass. The greater the volume fraction of particles,
the lower the overall energy storage. The optimal particle fraction was found to
depend on the ratio of the particle to PCM thermal conductivity, the thickness of
the PCM layer and the flowrate of the HTF within the tubes.
Another way to optimize the thermal response of a PCM heat exchanger dur-
ing charging and discharging is to design the exchanger to contain multiple PCMs
with different melting ranges. A multiple PCM exchanger for high temperature
applications was analyzed by Li et al. [47]. In this design, the PCM with the high-
est melt temperature is located at the inlet to the exchanger and occupies the first
1/3 of the shell length, the second PCM with a lower melt temperature fills the
middle third of the shell and the PCM with the lowest melt point fills the final 1/3
of the shell as shown in Fig. 2.10. In this way, as the HTF drops in temperature
as it traverses the tube, the temperature difference between the HTF and the melt
point of the PCM is adjusted by moving to lower melting point PCMs, maintain-
ing a more consistent heat flux.
While the analytical studies discussed above do not consider the effect of natu-
ral convection during the melt phase, experiments have shown that in practice the
26 2  Energy Storage Applications

Fig. 2.10  Cascaded
PCMs with different melt
temperatures

density differences between the solid and liquid PCM phases do induce significant
natural convection in the melt phase. This natural convection increases heat trans-
fer for a faster melt process. Akgun et al. [48] illustrate this experimentally, and
in fact show that this convective effect can be enhanced by inclining the shell 5°,
which decreases the melt time by 30 %.
Since the melting phase can be shortened using the effects of natural convec-
tion, it is the solidification process that is the limiting factor in the design of PCM
heat exchangers. During energy discharge, the PCM solidifies slowly on the cold
tubes. The heat transfer from PCM to HTF is conduction dominated and it takes
much longer to solidify the PCM than it did to melt it [49]. The energy discharge
(solidification) process must be shortened for effective implementation of PCM
heat exchangers. This can be done in several ways, including the use of fins or
embedded high conductivity particles to improve the thermal response. A long spi-
ral fin wrapped around the inner tube is used in a concentric tube heat exchanger
by Liu et al. [50]. The fin is shown to reduce the solidification time by up to 50 %.
Sanusi et al. [49] showed graphite nanofibers in paraffin PCM to be even more
effective with a decrease in solidification time by over 60 %.
It can be seen from this overview that PCMs can be implemented in heat
exchanger designs for quick charge and discharge of thermal energy to/from a
flowing heat transfer fluid. The PCM is typically installed on the shell side of a
shell and tube exchanger and the PCM used will depend on the temperature of the
application. The main design concern with PCM implementation in this manner
is the reduction of thermal resistance during both melting and solidification. The
thermal resistance is often reduced by the use of both internal and external fins on
the tubes, and by the use of high conductivity particles in the PCM itself.

2.7 Textile Designs

A relatively new, but extremely interesting, application of PCMs is in the design of


clothing that enhances comfort in extreme hot or cold conditions. In many ways,
this application is similar to the use of PCMs in building materials for mainte-
nance of comfortable environmental conditions, but with the implementation of
2.7  Textile Designs 27

PCMs in textiles rather than in concrete or wallboard. This can allow the design of
portable personal environmental control systems for both indoor and outdoor use.
By embedding PCMs within various textiles, both clothing and accessories with
thermal management capabilities can be designed. A handful of products have
been or are commercially available and include thermally regulating parkas, vests,
snow pants, underwear, socks, sleeping bags, blankets, duvets, mattresses and pil-
lowcases [51].
For hot weather gear, these products are designed to absorb excess heat from
the environment, thermally isolating the human body from excess temperatures. In
extreme climates, the heat is absorbed directly into the PCM which then maintains
a comfortable microclimate next to the body as the PCM melts. The goal, simi-
lar to that previously examined for thermal management of electronics, is to delay
the onset of steady-state temperatures as along as possible by matching the melt
time of the PCM material to the time of exposure to the high temperatures. Once
completely melted, the liquid PCM will rise in temperature until it reaches steady-
state, and must be cooled and solidified before its next use. If the useful melt time
can be extended to several hours, these types of garments can be extremely effec-
tive. In fact, the US military has shown significant interest in PCM textiles designs
over the years, funding multiple studies on gear design for use by troops deployed
to extreme environments. This type of system can also be effective in sports and
exercise gear.
For cold weather gear, the idea is similar but with the heat source reversed. In
this case, the PCM absorbs heat from the human body and as it melts, isolates the
body from extreme cold. The longer that the PCM can be maintained in melt phase
as it absorbs heat from the body, the longer the gear remains effective. This can be
particularly useful in certain types of parkas, including ski gear. As the user exer-
cises outdoors in cold weather the heat generated by the activity level is absorbed
into the melting of the PCM. During periodic rest breaks, such as on a ski lift, the
PCM begins to solidify, but maintains a constant temperature at the melt point,
increasing comfort levels for the user in comparison to normal gear which cools
down quickly when the user is at rest.
For textile based applications such as these, PCMs with melt temperatures from
12 to 35 °C will be most effective. Melt temperatures for cold weather gear should
be at the lower end of this range in order to ensure that melting is induced from the
body heat and activity levels. Even a 12 °C melt point can enhance user comfort in
outdoor winter environments of 0 °C and lower. Higher melt temperatures should
be selected for warm weather apparel to increase the length of the melt process
by minimizing the temperature differential between the environment and the melt
temperature. The various types of paraffin waxes are popular for textiles applica-
tions due to their suitable melt temperature ranges, their high latent heat and their
nontoxic, chemically inert nature [51]. Paraffin can be incorporated into the tex-
tiles in several different ways including the insertion of PCM-containing packets
in the design of the garment, in the application of coatings containing microen-
capsulated PCMs to the fabric, or in the direct incorporation of the PCM into the
fibers themselves.
28 2  Energy Storage Applications

The simplest, although perhaps least comfortable, method of incorporating


PCMS into garments is through the use of macroscale encapsulation of PCM in
flexible packs inserted into pockets in the garment itself. In one application, blis-
ter packs containing PCM were created which featured multiple cavities 10 mm in
diameter and 4 mm deep sealed with foil. The material was then used to line pro-
tective clothing for workers exposed at regular intervals to the heat from industrial
ovens [52]. Samples were made using four different PCMs in the blister packs and
the gear was tested in exposure to a 1500 W/m2 radiation heat flux at a distance of
50 cm. The testing occurred over three phases. In phase one the sample was pro-
tected from direct irradiation by a fireclay for two minutes. The fireclay was then
removed and the sample exposed to direct irradiation for 10 min. In the final phase
the fireclay was replaced for an additional 15 min of exposure time. All of the sam-
ples reached 120 °C on the front face of the gear by the end of 10 min of direct irra-
diation, but the PCM blister packs reduced the rate of temperature rise on the rear
of the sample, corresponding to the human-clothing interface, indicating significant
comfort improvements. The best design reached only 75 °C on the rear side after a
full 10 min of direct exposure, a reduction of 37 % in operating temperature.
Although effective for high heat flux exposure, for more moderate environ-
ments the use of large PCM lining packs is unnecessarily heavy and bulky.
Instead, the PCM can be contained within microscale polymer beads, and coated
onto the fabric as shown in Fig. 2.11. For instance, a fabric made by the Outlast
company and referred to as “Outlast/silk” composed of 57 % Outlast viscose (weft
yarn, 50 s) and 43 % silk (warp yarn, 22D) was coated with a mixture of 15 wt%
microencapsulated PCM, 15 wt% polyurethane binder and thickener (12 g/l) [53].
Thermal testing of the prepared fabrics showed that the microcapsules evenly
coated the fabric without affecting the airflow through the fabric, which is impor-
tant for breathability and comfort of the garments. The fabric coated with the PCM
capsules exhibited an energy storage capability of 7.71 J/g, a 95 % increase over
the base fabric [53].

Fig. 2.11  Fabric coated with


microbeads of PCM
2.7  Textile Designs 29

A study of the effectiveness of microencapsulated PCM coated fabric using a


human-clothing environment simulator (HCE) was completed by Yoo et al. [54]
for cotton coated with 8–30 μm microcapsules beads containing nonadecane PCM
with a melt temperature of 29 °C. The fabric was treated with the prepared micro-
capsules in an acrylic binder [54]. The materials were tested under exposure to a
hot temperature source of 34 °C and a low temperature condition of 10 °C. Sample
garments were comprised of four layers of material which varied parametrically
with the number and location of PCM treated layers. The PCM coated material
was shown to both reduce temperature rise when exposed to the hot conditions and
to reduce heat loss when exposed to cold conditions. The more layers coated with
PCM, the greater the thermal control effects within the garment [54]. The possible
disadvantages to coating treatments are concerns about fabric stiffness and about
the long term durability of the coating during repeated washings.
To eliminate the concerns with coating durability, fabrics can also be created
by incorporation of the PCM directly into the fibers themselves. For example, fib-
ers can be fabricated using a co-axial electrospinning technique which creates a
hollow fiber with PCM in the central core, such as those fabricated by Chen et al.
[55] with a cellulose acetate shell and a polyethylene PCM core. The fibers were
extruded using a stainless steel needle with an outer diameter of 0.9 mm and an
inner diameter of 0.6 mm. Thermal testing of the fibers showed good stability over
repeated cycling, and latent heat storage of 52–60 J/g during melting depending
on the exact composition of the fibers. However, the incorporation of the PCM
into the fiber sheath reduced the overall tensile strength of the fibers, which must
be considered in the application of the material. While this technique is still under
development, and few if any applications currently exist, the potential for the tech-
nology seems high.
It can be seen from this overview that PCMs can be effective in moderating
environmental temperature effects when embedded in textiles for use in apparel
and accessories design. The PCMs can be embedded by macroencapsulation in
large packs, by fabric coating with microencapsulated spheres and by direct con-
tainment within the fiber sheath. The PCM melt temperature range should be
within 11–35 °C and several commercial products are currently available. The bar-
riers to greater adoption seem to be comfort and durability issues, as well as cost
for the more complicated fabrication techniques.

2.8 Space Systems

The use of PCMs in space systems is one of the oldest and most enduring applica-
tions of these materials. The excessive temperature swings inherent in most space
applications demand a high level of thermal control and the passive and reliable
behavior of PCMs makes them a natural fit for these systems. In many early space
systems, PCMs were used to maintain a consistent temperature over time, and for
this reason NASA referred to their PCM designs as thermal capacitors. In many
30 2  Energy Storage Applications

Fig. 2.12  A thermal capacitor device

NASA applications the “thermal capacitors” were comprised of an external hous-


ing, and a finned inner area filled with a PCM as seen in Fig. 2.12 [1].
Thermal capacitors of this type, using paraffin as the PCM, were used on
Apollo 15, 16 and 17 for thermal control of various electronic components on
the Lunar Rovers [1]. Even more were used in the design of Skylab, the original
manned space station. Five were located in cooling systems, with two of these on
the primary Skylab space radiator fluid outlet, and three at the outlet of a smaller
radiator used to refrigerate food and biological waste samples. Ten other thermal
capacitors were mounted in the bottom of trays used to store human urine samples
[1] during transport from the Skylab Vehicle to earth, perhaps not the most glam-
orous application, but certainly functional.
The 1977 NASA tech brief “A Design Handbook for Phase Change Thermal
Control and Energy Storage Devices” [2] was one of the first comprehensive PCM
references, and with its property database and basic design guidelines, it is still
widely cited and used today. The use of PCM in the 0-g environment of space does
lead to certain design issues that do not occur in normal terrestrial applications. The
most serious design issue is the complete elimination of the buoyancy forces which
create natural convection in the melt phase. In the absence of standard natural con-
vection, Marangoni convection, which is driven by surface tension forces, can play
a more significant role in heat dissipation [2]. However, the overall melting rates
will be reduced. Additionally the volume contraction of the PCM which occurs
during solidification occurs slightly differently in space. In earth gravity, when
the PCM solidifies and contracts it always leaves an empty space at the top of the
system. However, in 0-g the location of the void is not at the top of the container.
Instead the void may change location, may be distributed as a number of smaller
voids throughout the system or may occur at the very center of the PCM mass [1].
In 2003, a review of thermal control solutions for contemporary space applica-
tions was published by Swanson and Birur [56]. In this paper a number of systems
are reviewed, and PCM is identified as the solution of choice for several systems
including the Mars rovers, Spirit and Opportunity, which were launched in 2003.
2.8  Space Systems 31

In these systems, paraffin waxes, dodecane and hexadecane, were used in thermal
storage capsules to control battery temperature and damp out temperature fluctua-
tions. Interestingly, this paper also describes the development of a PCM actuated
thermal switch in which the volume change of the wax under heating or cooling
would make or break a contact actuated switch that prevents warm objects from
getting too cold, or cold devices from getting too warm in extreme environments.
In an application somewhat similar to the human urine transport from Skylab,
PCM is currently being used for thermal stabilization of biological samples, such
as cell samples, used in scientific testing aboard the International Space Station
(ISS) during transport to and from the earth. PCM based sample containers were
developed by the European Space Agency (ESA) and are referred to as ECCOs
(ESA Conditioned Containers). The ECCOs utilize containers of two different
sizes matched with modular PCM cartridges with different melt points depending
on the application. The thermal control points available include below −20 °C,
from 2–10 °C or above 23 °C [57]. The different container sizes allow transport
on the Space Shuttle, Soyuz, Progress (an older Russian transport vehicle), ATV
(the EAS vehicle) and HTV (Japanese transport vehicle), and is compatible with
the new SpaceX dragon vehicle. Another smaller container (Mini-ECCO) can be
transported in the Soyuz crew re-entry capsule. In operation, the containers are
“pre-charged” by fully solidifying the PCM prior to launch so that the PCM melts
slowly from supercooled conditions through trip, maintain the desired tempera-
ture. These transport containers have been used successfully since 2008 [57].
In the reverse of this application, premelted PCM is used to provide heat to sys-
tems onboard launch vehicles during the trip from earth to the ISS [58, 59]. While
in most cases, the transport vehicle can provide heating power during the actual
trip, during the time that it takes for payload transition at the ISS, the transport vehi-
cle can be unpowered for as long as 6 h. This extensive length of time without heat
in space can lead to the payload temperature falling below the electronics cold sur-
vival limit of −30 °C. By melting the PCM prior to power cutoff, enough energy
can be stored for release into the payload to keep it warm for several hours [58].
Choi [59] applied this concept model to the design of a particular application,
the transport of the Neutron Star Interior Composition Explorer (NICER) in the
SpaceX dragon transport vehicle. The payload volume of the SpaceX vehicle is
3.6 m in diameter by 2.3 m tall with the interior insulated with a 7.62 cm thick
acoustic blanket [59]. NICER is an X-ray timing and spectroscopy instrument to
be used to study neutron star emissions from aboard the ISS. NICER is due to be
launched in Dec. 2016. NICER features a 1.18 m by 0.87 m by 0.82 m Instrument
Optical Bench (IOB) with the spectroscopy instruments. The electronics to run the
instruments are all mounted to a single heat spreader which is thermally isolated
from the IOB.
Several paraffins were considered for use as the PCM, including dodecane, tet-
radecane, and tridecane. Ultimately dodecane, which features a high latent heat of
fusion (217 kJ/kg) and a low melting point (−10 °C) was chosen. The proposed
design features four PCM panels located between the electronics heat spreader and
32 2  Energy Storage Applications

the IOB, each with 0.51 kg of paraffin. Six additional paraffin panels are used to
wrap around the IOB, each with 0.38 kg of PCM. The 4.32 kg of dodecane will
store up to 937 kJ of thermal energy as latent heat.
The model assumes that NICER is removed from the Dragon trunk at orbit
noon and immediately begins cooling down. When the payload reaches −10 °C,
the PCM solidifies and then will subcool after full solidification. After 6 h with-
out power, the model predicts a final payload temperature of −13 °C, which is
adequate protection from the electronics failure point of −30 °C.
As this overview illustrates, PCMs have a long and successful history in many
varied space applications from vehicle electronics to sample transport. The PCMs
used in these systems are most often paraffin, and feature low melt points, as low
as −10 °C, to protect systems from the harsh space environment.

References

1. Humphries WR (1974) Performance of finned thermal capacitors. NASA technical note


D-7690
2. Humphries WR, Griggs EI (1977) A design handbook for phase change thermal control and
energy storage devices. NASA technical paper 1074
3. Theunissen PH, Buchlin JM (1983) Numerical optimization of a solar air heating system
based on encapsulated PCM storage. Sol Energy 31:271–277
4. Wang JCY, Lin S, Kwok CCK, Vatistas GH (1984) An analytical study of heat exchanger
effectiveness and thermal performance in a solar energy storage system with PCM. J Sol
Energy Eng 106:231–233
5. Santamouris MJ, Lefas CC (1988) On the coupling of PCM stores to active solar systems. Int
J Energy Res 12:603–610
6. Solomon AD (1979) Design criteria in PCM wall thermal storage. Energy 4:701–709
7. Grodzka P, Price J, Serbin C, Solomon A (1982) On the development of heat storage building
materials. In: Proceedings of the 20th intersociety energy conversion engineering conference,
vol 4, pp 2070–2073
8. Bathel AG, Viskanta R (1980) Heat transfer at the solid-liquid interface during melting from
a horizontal cylinder. Int J Heat Mass Transf 23:503–1493
9. Ho CJ, Viskanta R (1984) Heat transfer during melting from an isothermal vertical wall.
J Heat Transf 106:12–19
10. Sparrow EM, Zumbrunne ML (1986) In-tube melting in the presence of circumferentially
nonuniform heating. Int J Heat Mass Transf 29:1629–1637
11. Sparrow EM, Geiger GT (1986) Melting in a horizontal tube with the solid either constrained
or free to fall under gravity. Int J Heat Mass Transf 29:1007–1019
12. O’Connor JP, Weber RM (1997) Thermal management of electronic packages using solid-to-
liquid phase change techniques. Int J Microcircuits Electron Packag 20:593–601
13. Pal D, Joshi YK (1997) Application of phase change materials to thermal control of elec-
tronic modules: a computational study. J Electron Packag 119:40–50
14. Vesligaj MJ, Amon CH (1999) Transient thermal management of temperature fluctuations
during time varying workloads on portable electronics. IEEE Trans Compon Packag Technol
22:541–550
15. Sarier N, Onder E (2012) Organic phase change materials and their textile applications: an
overview. Thermochim Acta 540:7–60
16. Krishnan S, Garimella SV (2004) Thermal management of transient power spikes in elec-
tronics—phase change energy storage or copper heat sinks? J Electron Packag 126:308–316
References 33

17. Hodes M, Weinstein RD, Pence SJ, Piccini JM, Manzione L, Chen C (2002) Transient
thermal management of a handset using phase change material (PCM). J Electron Packag
124:419–426
18. Setoh G, Tan FL, Fok SC (2010) Experimental studies on the use of a phase change material
for cooling mobile phones. Int Comm Heat Mass Transf 37:1403–1410
19. Lingamneni S, Ashedgi M, Goodsen KE (2014) A parametric study of microporous metal
matrix phase change material composite heat spreaders for transient thermal applications.
In: IEEE intersociety conference on thermal and thermomechanical phenomena in electronic
systems
20. Ling Z, Chen J, Fang X, Zhang Z, Xu T, Gao X, Wang S (2014) Experimental and numeri-
cal investigation of the application of phase change materials in a simulative power batteries
thermal management system. Appl Energy 121:104–113
21. Feldman D, Banu D, Hawes DW (1995) Development and application of organic phase
change mixtures in thermal storage gypsum wallboard. Sol Energy Mater Sol Cells
36:147–157
22. Athienitis AK, Liu C, Hawes D, Banu D, Feldman D (1997) Investigation of the ther-
mal performance of a passive solar test room with wall latent heat storage. Build Environ
32:405–410
23. Zhang H, Xu Q, Zhao Z, Zhang J, Sun Y, Sun L, Xu F, Sawada Y (2012) Preparation and
thermal performance of gypsum boards incorporated with microencapsulated phase change
materials for thermal regulation. Sol Energy Mater Sol Cells 102:93–102
24. Tyagi VV, Kaushik SC, Tyagi SK, Akiyama T (2011) Development of phase change materi-
als based microencapsulated technology for buildings: a review. Renew Sustain Energy Rev
15:1373–1391
25. Parameshwaran R, Kalaiselvam S, Harikrishnan S, Elayaperumal A (2012) Sustainable
thermal energy storage technologies for buildings: a review. Renew Sustain Energy Rev
6:2394–2433
26. Han B, Zhang K, Yu X (2013) Enhance the thermal storage of cement based composites with
phase change materials and carbon nanotubes. J Sol Energy Eng 135:024505
27. Kosny J, Biswa K, Miller W, Kriner S (2012) Field thermal performance of naturally venti-
lated solar roof with PCM heat sink. Sol Energy 86:2504–2514
28. Barlev D, Vidu R, Stroeve P (2011) Innovation in concentrated solar power. Sol Energy
Mater Sol Cells 95:2703–2725
29. Denholm P, Mehos M (2011) Enabling greater penetration of solar power via the use of csp
with thermal energy storage. NREL technical report NREL/TP-6A20–52978
30. Sioshansi R, Denholm P (2010) The value of concentrating solar power and thermal energy
storage. IEEE Trans Sustain Energy 1:73–183
31. Gil A, Medrano M, Martorell I, Lázaro A, Dolado P, Zalba B, Cabeza LF (2010) State of the
art on high temperature thermal energy storage for power generation. Part 1-Concepts, mate-
rials and modellization. Renew Sustain Energy Rev 14:31–55
32. Adinberg R, Zvegilsky D, Epstein M (2010) Heat transfer efficient thermal storage for steam
generation. Energy Convers Manag 51:9–15
33. Robak CW, Bergman TL, Faghri A (2011) Economic evaluation of latent heat thermal energy
storage using embedded thermosyphons for concentrating solar power applications. Sol
Energy 85:2461–2473
34. Nithyanandam K, Pitchumani R (2014) Cost and performance analysis of concentrating solar
power systems with integrated latent thermal energy storage. Energy 64:793–810
35. Sole C, Medrano M, Castell A, Nogues M, Mehling H, Cabeza LF (2008) Energetic and
exergetic analysis of a domestic water tank with phase change materials. Int J Energy Res
32:204–214
36. Kenisarin M, Mahkamov M (2007) Solar energy storage using PCMs. Renew Sustain Energy
Rev 11:1913–1965
34 2  Energy Storage Applications

37. Ozturk HH (2005) Experimental evaluation of energy and exergy efficiency of a seasonal
latent heat storage system for greenhouse heating. Energy Convers Manag 46:1523–1542
38. Esen M, Durmus A, Durmusm A (1998) Geometric design of solar-aided latent heat store
depending on various parameters and phase change material. Sol Energy 62:19–28
39. Esen M (2000) Thermal performance of a solar aided latent heat store used for space heating
by heat pump. Sol Energy 69:15–25
40. Benli H (2011) Energetic performance analysis of a ground-source heat pump system with
latent heat storage for a greenhouse heating. Energy Convers Manag 52:581–589
41. Arkar C, Medved S (2005) Influence of accuracy of thermal property data of a phase change
material on the result of a numerical model of a packed bed latent heat storage with sphere.
Thermochim Acta 438:192–201
42. Lee YT, Hong SW, Chung JD (2014) Effects of capsule conduction and capsule outside
convection on the thermal storage performance of encapsulated thermal storage tanks. Sol
Energy 110:56–63
43. Narasimhan N, Bharath R, Ramji S, Tarun M, Arumugam A (2014) Numerical studies on the
performance enhancement of an encapsulated thermal storage unit. Int J Heat Mass Transf
84:184–195
44. Jegadheeswaran S, Pohekar SD (2009) Performance enhancement in latent heat thermal stor-
age system: a review. Renew Sustain Energy Rev 13:2225–2244
45. Al-abibi A, Mat SB, Sopian K, Sulaiman MY, Lim CH, Abdulrahman T (2010) Review
of thermal energy storage for air conditioning systems. Renew Sustain Energy Rev
16:5802–5819
46. Zhang Y, Faghri A (1996) Heat transfer enhancement in latent heat thermal energy storage
system by using the internally finned tube. Int J Heat Mass Transf 39:73–3165
47. Li YQ, He YL, Song HJ, Xu C, Wang WW (2013) Numerical analysis and parameters opti-
mization of shell and tube heat storage unit using three phase change materials. Renew
Energy 559:92–99
48. Akgun M, Aydin O, Kaygusuz K (2007) Experimental study on melting/solidification charac-
teristics of a paraffin as PCM. Energy Convers Manag 28:669–678
49. Sanusi O, Warzoha R, Fleischer AS (2011) Energy storage and solidification enhancements
of phase change materials embedded with nanofibers. Int J Heat Mass Transf 54:4429–4436
50. Liu Z, Sun X, Ma C (2005) Experimental study of the characteristics of solidification of
stearic acid in an annulus and its thermal conductivity enhancement. Energy Convers Manag
46:971–984
51. Sarier N, Onder E (2012) Organic phase change materials and their textile applications: an
overview. Thermochim Acta 540:7–60
52. Buhler M, Popa AM, Scherer LJ, Lehmeier FKS, Rossi RM (2013) Heat protection by differ-
ent phase change materials. Appl Therm Eng 54:64–359
53. Yan C, Yu Z, Yang B (2013) Improvement of thermoregulating performance for outlast/silk
fabric by the incorporation of polyurethane microcapsule containing paraffin. Fibers Polym
14:1290–1294
54. Yoo H, Lim J, Kim E (2013) Effects of the number and position of phase change material
treated fabrics on the thermo-regulating properties of phase change material garments. Text
Res J 83:671–682
55. Chen C, Zhao Y, Liu W (2013) Electrospun polyethylene glycol/cellulose acetate phase
change fibers with core-sheath structure for thermal energy storage. Renew Energy
60:222–225
56. Swanson TD, Birur GC (2003) NASA thermal control technologies for robotic spacecraft.
Appl Therm Eng 23:1055–1065
57. Neri G, Koehler A, De Parolis M, Zolesi V (2012) ESA conditioned container: a system for
passive temperature controlled transportation of experiments for the international space sta-
tion. In: Proceedings of the international astronautical congress IAC 3947–3954
References 35

58. Choi MK (2012) Using premelted phase change material to keep payload warm without
power for hours in space. In: Proceedings of 10th international energy conversion engineer-
ing conference AIAA 2012–3894
59. Choi MK (2013) Using paraffin with −10 °C to 10 °C melting point for payload thermal
energy storage in SpaceX dragon trunk.In: Proceedings of 11th international energy conver-
sion engineering conference
Chapter 3
Types of PCMs and Their Selection

3.1 Important PCM Material Properties

There are many different types of PCMs available, but the vast majority fall into
three main classifications: organics, inorganics, and liquid metals. This chapter is
devoted to looking at each of these main types in depth. However, there are sev-
eral material selection considerations that are common to all PCM applications,
regardless of primary material types, so we will explore those first.
First and foremost, the melting point is the property of primary consideration
when choosing a material. An obvious consideration is that the melting point of
the PCM must be both below the temperature of the heat source and above the
ambient conditions to which the device will be exposed. Certainly, it does not
make sense to choose a material which will already be in liquid phase before
absorbing heat from the intended source. When choosing the operating point, one
must also be certain to consider all the extremes in temperature to which your
device will be subjected. For instance, in designing PCMs for portable electron-
ics—what are the highest possible ambient temperatures to which your device may
be exposed? Will the PCM be as effective in Cairo as in Helsinki if you have a
global market?
For systems that are designed preferentially for thermal management (mainte-
nance of a specific operating point), the general guide is to pick a PCM with the
highest possible melt point that is still below the desired thermal control point. For
instance, if the desired operating point of a portable electronic system is 80 °C,
then it will be advantageous to select a PCM with a melt temperature range around
70–75 °C. The advantage to using a high melting point material in these situa-
tions is that it leads to a longer melting time, and thus a longer effective thermal
management duty cycle before the PCM is completely melted. Systems with a
higher melting point will dissipate more heat to the ambient while simultaneously

© The Author(s) 2015 37


A.S. Fleischer, Thermal Energy Storage Using Phase Change Materials,
SpringerBriefs in Thermal Engineering and Applied Science,
DOI 10.1007/978-3-319-20922-7_3
38 3  Types of PCMs and Their Selection

melting, due to the higher differential from the system to the ambient, thus extend-
ing the melt time. A study by Leoni and Amon [1] illustrates the effect of PCM
melt temperature by comparing the transient absorption time for two PCMs with
similar latent heats, but different melt temperatures. Changing the selected PCM
from one with a melt temperature of 36 °C to one with a melt temperature of
44 °C, with all other system variables held constant, extended the transient melting
time, and thus the effective operation of the system, by 50 %.
Once you have narrowed in on a desirable melt range for the PCM, the second
most important criterion is generally the latent heat of fusion of the material. The
latent heat is the measure of how much energy can be stored in a specified mass of
material (kJ/kg) during melting and released during solidification. The higher the
latent heat of fusion, the more energy efficient the system is, and the less mass can
be used, minimizing the size of the system. You want to select the material with
the highest latent heat of fusion with a melt temperature in the range previously
determined.
Materials with high latent heats of fusion also tend to have relatively high
specific heats. Specific heat refers to the amount of energy it takes to raise one
kg of mass of the material one degree kelvin and is reported in units of J/kg K.
For energy storage applications, having a high specific heat is as advantageous,
as having a high latent heat of fusion. This is because a significant amount of
energy can then be stored during both the sensible heating stage (as the PCM
approaches the melt point) and during the latent heating phase (through the melt
transition).
There are several other criteria that should be considered when selecting a
PCM material. The material should exhibit stability, both chemically and physi-
cally over repeated thermal cycling with repeatable and consistent melting/freez-
ing cycles. As will be seen later in this chapter, certain PCMs may break down
over extended cycling, so one must be aware of this possibility. The compatibil-
ity of the selected material with the casing and other materials should also be
considered as certain PCMs tend to be corrosive and/or chemically incompati-
ble. It is also preferred that the PCM be environmentally safe, nonflammable and
nontoxic.
Many PCMs exhibit a density different between the solid and liquid phase,
which in most cases leads to contraction upon solidification. This density differ-
ence must be considered in the design of any containment structure.
Finally, it is desirable for the PCM to have a high thermal conductivity to
prevent thermal bottlenecking at the source, although it is challenging to find
materials that have both high latent and specific heats as well as high thermal con-
ductivities. In most cases, the desire to store more energy (high latent heat) will
win out over the need for high thermal conductivity simply due to the nature of
the applications. Methods to address the low thermal conductivity of most PCMs
will be discussed in in Chap. 4. From the above discussion it can be seen that the
thermal properties of greatest interest in PCMs are the melt temperature, the latent
heat, specific heat, and thermal conductivity.
3.2  Organic PCMs 39

3.2 Organic PCMs

Organics are arguably the most popular type of PCM. Organics can include a wide
range of PCMs such as those in the alkane (paraffin) family (CnH2n+2), and the
fatty acids family (CH3(CH2)2nCOOH). The organic PCMs tend to be abundantly
available, relatively inexpensive and easy to work with. As seen in Fig. 3.1, the
alkanes and the fatty acids have similar physical characteristics, exhibiting a white
surface which is soft and waxy in appearance.
The organics, specifically paraffin, are the most commonly studied PCMs for
electronics thermal management. The melt range of materials in the paraffin fam-
ily tend to be from 35 °C to about 70 °C, depending on the specific hydrocarbon
structure, making them an excellent match with electronics which have a maxi-
mum junction temperature of around 85 °C. Many researchers, including many
of those discussed in Sect. 2.1 are using paraffin based PCMs for thermal man-
agement of electronics and there is a significant amount of ongoing work to fully
classify and characterize the thermal properties of different paraffins in both the
solid and liquid phases to enable their effective use in systems [2–5]. Despite their
popularity, it remains challenging to find accurate thermal property data for the
different organics, particularly when the materials are in the liquid phase.
Fatty acids have lower melting points than paraffins, making them better suited
for applications related to human comfort. For example, mixtures of fatty acids
such as stearic acid, lauric acid and capric acid have been embedded in mats of
polyacrylonitrile fibers [6] and in diatomite, perlite and vermiculite [7]. In these
cases the fatty acid PCM composite materials were intended for use as energy
storage building materials with the aim to reduce HVAC costs.
For organic PCMs, it is typically more accurate to speak of a melt range,
rather than a melt temperature. A specific melt temperature implies a sharp gra-
dient in behavior, whereas for most organic PCMs, the melt occurs more slowly
over a range of temperatures. This can be seen most clearly considering a tran-
sient heating curve. Figure 3.2 shows a heating curve resulting from the use of the

Fig. 3.1  Paraffin wax (left) and stearic (fatty) acid (right)


40 3  Types of PCMs and Their Selection

Fig. 3.2  Example DSC curve for paraffin

differential scanning calorimetry (DSC) technique to measure an organic PCM’s


specific heat, latent heat and melt range. Differential scanning calorimetry meas-
ures the changes in the heat capacity of a material as temperature increases. This
allows one to determine both the specific heat of the material by tracking changes
in capacity when the material is purely solid or purely liquid, and also the latent
heat of the material as it transitions from solid to liquid over the melt range. The
melt range is commonly known as the “mushy zone” during which the PCM first
softens and then melts. The melt range can be seen in Fig. 3.2. The mushy zone
begins as the heating curve begins to dip around 319 K and peaks at the peak melt
point of 331 K. The mushy zone for paraffins is wider than for fatty acids, and can
be several degrees. Blended paraffins, like those commonly available from com-
mercial suppliers, can have mushy zones which extend 10–20 °C as seen here.
Paraffins also commonly exhibit a solid-solid phase transition prior to melting,
which can also be seen in Fig. 3.2.
There are many advantages to using the organic PCMs. Chief among these are
melt temperatures that match well with the electronics thermal management and
building systems management applications. Organics also have high latent heats of
allowing them to store a high amount of energy in a smaller mass. Common paraf-
fins exhibit latent heats in the range of 200–300 kJ/kg and fatty acids in the range
of 100–200 kJ kg.
Considering the other criteria mentioned earlier this chapter for PCM selection,
organics are chemically and physically stable and will maintain their performance
over repeated thermal cycling without breaking down or separating. Organics are
compatible with a wide range of materials and do not act to corrode or otherwise
attack most common casing materials. Organics are flammable due to their hydro-
carbon nature, but have flash points close to 200 °C, well outside the operating
range of their common applications.
The main disadvantage to most organics is that they exhibit extremely low
thermal conductivities. The low thermal conductivity (for example ~0.2 W/m K
for most paraffin waxes) decreases their effectiveness in both energy storage and
3.2  Organic PCMs 41

Table 3.1  Thermal properties of some common organic PCMs


Name Type Tm Latent heat ρ (kg/m3) Cp (kJ/kg) k (W/m K) References
(°C) (kJ/kg)
Octadecane Paraffin 29 244 814 (sol) 2150 (sol) 0.358 (sol) [10]
724 (liq) 2180 (liq) 0.152 (liq)
Heneicosane Paraffin 41 294.9 773 (liq) 2386 (liq) 0.145 (liq) [2, 8]
Tricosane Paraffin 48.4 302.5 777.6 (liq) 2181 (liq) 0.124 (liq) [2, 8]
Tetracosane Paraffin 51.5 207.7 773.6 (liq) 2924 (liq) 0.137 (liq) [2, 8]
IGI 1230A Blended 54.2 278.2 880 (sol) 2800 (liq) 0.25 (sol) [3, 9]
paraffin 770 (liq) 0.135 (liq)
Oleic acid Fatty acid 13 75.5 871 (liq) 1744 (liq) 0.103 (liq) [2, 12, 13]
Capric acid Fatty acid 32 153 1004 (sol) 1950 (sol) 0.153 (liq) [11, 12]
878 (liq) 1720 (liq)
Lauric acid Fatty acid 44 178 1007 (sol) 1760 (sol) 0.147 (liq) [11, 12]
965 (liq) 2270 (liq)
Palmitic acid Fatty acid 64 185 989 (sol) 2200 (sol) 0.162 (liq) [11, 12]
850 (liq) 2480 (liq)
Stearic acid Fatty acid 69 202 965 (sol) 2830 (sol) 0.172 (liq) [11, 12]
848 (liq) 2380 (liq)

thermal management applications. The low thermal conductivity creates a high


thermal resistance to heat flow preventing the heat from effectively penetrat-
ing into the PCM and initiating the melt process. This can lead to isolation of the
melt process near the heat source, resulting in a superheated liquid layer at the
source and a solid layer farther from the source. Therefore, much research work
has focused on addressing the challenge of increasing the thermal diffusivity and
thermal conductivity of organic phase change materials by using embedded high
conductivity structures of various designs. This will be discussed in Chap. 4.
The properties of several common organic PCMs can be found in Table 3.1.
There are several comprehensive reviews available that list the available proper-
ties of many more possible organic PCMs and combinations of organics PCMs
[11–13]. In some cases, using blended organic PCMs allows customized melt
temperature ranges and specific thermal properties which can be tailored to the
specific application. Thermal engineers are urged to treat these reported values as
guidelines as it becomes clear from a review of the literature that in many cases
there are conflicting values of organic PCM thermal properties reported and the
discrepancies as be as high as 10–50 % from source to source [12].

3.3 Inorganic PCMs

The family of inorganic PCMs includes the salts and salt hydrates. Salt hydrates
are combinations of members of the inorganic salt family (oxides, carbonates, sul-
fates, nitrates and halides) with water molecules in a specific ratio. Salt hydrates
are named using the name of the salt compound *n H2O. Common salts and salt
42 3  Types of PCMs and Their Selection

hydrates used as PCMs include MgCl2 · 6H2O, CaCl2 · 6H2O, Na2SO4 · 10H2O


(Glauber’s Salts), NaNO3, KNO3, KOH, MgCl2, and NaCl. Salts and salt hydrates
have similar characteristics. The salt hydrates exhibit a three dimensional struc-
ture which is open enough to allow water molecules to fit inside the crystal lattice.
Figure 3.3a shows the crystal structure of Calcium Chloride which easily absorbs
H2O to become Calcium chloride hexahydrate. The crystal structure is seen to be
quite open in comparison to the tightly packed crystal structure of NaCl shown in
Fig. 3.3b which is not easily hydrated. All of the salts and salt hydrates exhibit a
well-defined crystalline structure. Physical samples of salts and salt hydrates are
seen in Fig. 3.4.
Salts and salt hydrates can be found with melting points over a very wide range
of operating point from 10 to 900 °C. In the lower melt temperature ranges, where
the melting points overlap with those of the paraffins and fatty acids, salts and salt
hydrates are not commonly used. However, in the higher melt temperature ranges
the inorganics become the primary choice for PCMs. Inorganics are widely used in
solar energy applications where the melt temperatures are extremely high, such as
in concentrating solar plants, as discussed in Sect. 2.3.1. Currently the operating
concentrating solar plants are using inorganic salts entirely in the molten phase to
store energy using sensible heat. This is possible due to the high specific heat that
these salts and salt hydrates possess. However, much research is ongoing in the

Fig. 3.3  a Calcium chloride, b sodium chloride

Fig. 3.4  a Sodium chloride, b magnesium chloride hexahydrate


3.3  Inorganic PCMs 43

solar field in order to take advantage of the latent heat as well as the sensible heat
capability of these materials in order to store even higher amounts of energy for
use during times in which incident sunlight is not available [14–16]. The Andasol
plant in southern Spain stores heat in a molten mixture of 60 % sodium nitrate and
40 % potassium nitrate using a sensible heating system and went online in 2008.
Expansions were added in 2009 and 2011, and both exhibit molten salt storage.
Few of the organics are suitable for use in applications above 100 °C, so the
extended range of the inorganics (up to 900 °C) is noteworthy. The inorganics
have sharp transitions at the melt point, rather than exhibiting an extended mushy
region, have latent heats that are comparable to the organics, and have higher ther-
mal conductivities than the organics, although in many cases their thermal con-
ductivity still needs to be further enhanced. The inorganics also exhibit smaller
density changes as they transition from liquid to solid, reducing containment
issues. The inorganics have higher densities than the organics, which can be either
a positive or negative depending on the application. The higher density results in
a higher mass for a given volume, which can lead to greater energy density since
energy is stored on per unit mass basis. This is beneficial in most cases. However,
it can also lead to greater weights, which can be detrimental in some applications.
However, despite these advantages, in all applications except the high tem-
perature ranges where few options exist, the negative attributes of salts and salt
hydrates limit their implementation. The most significant issue with salt hydrates
is that they exhibit significant instabilities. The hydrates tend to dehydrate as the
water is driven off during the heating cycles, leading to breakdown of the material
itself, and all of the salts exhibit a tendency to break down and separate into con-
stituent parts over repeated cycling Several studies have examined the instability
of these materials and note their limited application at this point [17, 18].
The inorganics tend to aggressively attack many common casing materials
making their design implementation challenging. Thermal engineers are cautioned
to check all compatibilities careful before working with the salts and salt hydrates.
Despite their corrosiveness however, most of the inorganic salts are not highly
flammable or combustible.
Finally, the inorganics tend to exhibit a high propensity towards supercooling,
meaning that they do not begin to freeze at their solidification point, but instead
can require substantial subcooling before solidification nucleation begins. This
is a form of undesirable instability that makes it difficult to solidify the melted
PCM for the next thermal cycle. This can be solved to a degree by using nucle-
ating agents with a selected PCM. The nucleation agents provide a surface from
which the crystal formation can more readily occur, greatly reducing the rate of
supercooling. For example NaCl at low seed rates (<1 %) can provide nucleation
initiation sites for CaCl2  · 6H2O. In one study, 1 wt% NaCl in calcium chloride
hexahydrate eliminated phase separation and improved supercooling over 1000
thermal cycles [19].
The properties of several common inorganic PCMs can be found in Table 3.2.
As with the organic PCMs there are comprehensive reviews available that list the
available properties of many more possible organic PCMs and combinations of
44 3  Types of PCMs and Their Selection

Table 3.2  Thermal properties of common inorganic PCMs


Chemical Name Peak melt Latent Density Thermal References
formula point (°C) heat (kg/m3) conductivity
(kJ/kg) (W/m K)
MgCl2 · 6H2O Magnesium 117 168.6 1450 (liq) 0.579 (liq) [11]
chloride 1569 (solid) 0.694 (solid)
hexahydrate
CaCl2 · 6H2O Calcium 29 170–192 1562 (liq) 0.561 (liq) [13]
chloride 1802 (solid) 1.008 (solid)
hexahydrate
NaSO4 · 10H2O Glauber’s 32 251 1485 (solid) 0.544 [13]
salts
NaNO3 Sodium 307 172 2260 (solid) 0.5 [11]
nitrate
KNO3 Potassium 333 266 2110 (solid) 0.5 [11]
nitrate
MgCl2 Magnesium 714 452 2140 (solid) NA [11]
chloride
NaCl Table salt 802 492 2160 (solid) 5.0 [11]
(sodium
chloride)
KF Potassium 857 452 2370 (solid) NA [11]
fluoride

organics PCMs [11–13] and thermal engineers are urged to treat these reported
values as guidelines as in many cases there are wide ranges of inorganic PCM
thermal properties reported [12].

3.4 Metal and Metal Alloy PCMs

The family of metal and metal alloys is perhaps the most underused of all the
common PCM families, perhaps due to the low latent heat that most of these mate-
rials exhibit. However, despite this the metals do show promise in certain appli-
cations. The metal and metal alloys PCMs include a number of materials with
melting points in the range of desired PCM applications. Many of these metals
are easy to work with and have been used for years in other applications requir-
ing molten metals, for example, the hobby of casting tin soldiers. Of course the
practice of using lead for many of these low melt point metal applications, such as
casting bullets, is now obsolete due to the proven adverse health effects of working
with lead. There do remain however, many other promising and easy to work with
soft metals which transition in the temperature ranges of interest here.
The metals that exhibit the most promise in low temperature applications are
Cesium, Gallium, Indium, Tin and Bismuth, while the metals for high temperature
3.4  Metal and Metal Alloy PCMs 45

applications include zinc, magnesium, aluminum and their alloys. While metals
are not actively used in many applications, they are starting to attract some inter-
est due to their high thermal conductivities and their physical and chemical stabil-
ity at high temperatures. Ge and Liu [20] consider the use of gallium as a PCM
thermal management solution for smart phones where the high thermal conductiv-
ity is advantageous for quick charge and discharge of the material. However, the
tendency of Gallium to exhibit a high degree of supercooling negates the benefits
of the high thermal conductivity to a certain extent. Application of the high melt
temperature metals in solar energy systems also should be considered due to the
significant stability issues of the inorganics salts as seen in Sect. 3.3.
The metals and metal alloys cover a wide range of melt temperatures, and
exhibit a much sharper melt transition than the organics. Metals such as Cesium
and Gallium can melt at ambient conditions on a warm day (28.65 and 29.8 °C
respectively), while Magnesium melts at 648 °C and Aluminum melts at 661 °C.
Custom alloys can be designed to create tailored melt points for specific applica-
tions. The metals exhibit excellent chemical and physical stability and are com-
patible with a wide range of casing materials, eliminating many of the negative
characteristics of the salts.
The metals are the only PCMs that do not suffer from low thermal conductivi-
ties, with values that are several orders of magnitude larger than those of the organ-
ics and the inorganics. Aluminum of course is well-known for its excellent thermal
conductivity (237 W/m K), and while the low melt point metals don’t approach
this value, they still exhibit thermal conductivities in the range of 8–40 W/m K, a
significant improvement over the 0.2 W/m K of the paraffin family. This charac-
teristic eliminates any concerns about the thermal resistance of the PCM itself and
significantly reduces any issues with thermal isolation at the heat source.
The main detriments to the metallic PCMs are their high densities and low
latent heats of fusion. The high densities do offset some of the low latent heat
concern, as high densities lead to greater energy densities. Larger masses in the
same volume will store more energy/volume than low masses. However, the latent
heats exhibited by the metals are low enough that the higher energy density can’t
entirely offset the low energy storage values, and typically greater masses and vol-
umes are required for metals, leading to weight concerns with the high densities.
The latent heats for the metals and metal alloys can be an order of magnitude
lower than organics in the same melt temperature range. For instance, Cesium and
Gallium have latent heats of 16.4 and 80.1 kJ/kg respectively while Octadecane
has a latent heat of 244 kJ/kg at the same melting temperature. The higher melt
temperature metals are more competitive with the inorganic salts, exhibiting latent
heats of 300–500 kJ/kg, which is the same range exhibited by the high temperature
salts.
The properties of several common metallic PCMs can be found in Table 3.3.
Due to the limited application of metallic PCMs at this point, there are not as
many comprehensive reviews of these materials, but [21] may be useful.
46 3  Types of PCMs and Their Selection

Table 3.3  Thermal properties of common metallic PCMs


Name Tm (°C) Latent heat ρ (kg/m3) Cp (kJ/kg) k (W/m K) References
(kJ/kg)
Cesium 28.65 16.4 1796 0.236 17.4 [21]
Gallium 29.8 80.1 5907 0.237 29.4 [21]
Indium 156.8 28.59 7030 0.23 36.4 [21]
Tin 232 60.5 730 0.221 15.08 [21]
Bismuth 271.4 53.3 979 0.122 8.1 [21]
Zinc 419 112 7140 0.39 (sol) 116 [22]
0.48 (liq)
Al59-35Mg-6Zn 443 310 2380 1.63 (sol) NA [23]
1.46 (liq)
Al54-22Cu- 520 305 3140 1.51 (sol) NA [23]
18Mg-6Zn 1.13 (liq)
Al65-30Cu-5Si 571 422 2730 1.3 (sol) NA [23]
1.2 (liq)
Al88-Si12 576 560 2700 1.038 (sol) 160 [23]
1.741 (liq)
Mg 648 365 1740 1.27 (sol) 156 [22]
1.37 (liq)
Al 661 388 2700 0.9 (sol) 237 [22]
0.9 (liq)

References

1. Leoni N, Amon CH (1997) Transient thermal design of wearable computers with embedded
electronics using phase change materials. In: Proceedings 32nd national heat transfer confer-
ence, vol 5, pp 49–56
2. O’Connor W, Warzoha R, Weigand R, Fleischer AS, Wemhoff AP (2014) Thermal property
prediction and measurement of organic phase change materials in the liquid phase at the
melting point. Appl Energy 132:496–506
3. Warzoha R, Weigand R, Fleischer AS (2015) Temperature-dependent thermal properties of
a paraffin phase change material embedded with herringbone style graphite nanofibers. Appl
Energy 137:716–725
4. Warzoha R, Sanusi O, McManus B, Fleischer AS (2013) Development of methods to fully
saturate carbon foam with paraffin wax phase change material for energy storage. J Sol
Energy Eng 135:021007
5. Chintakrinda K, Warzoha R, Weinstein RD, Fleischer AS (2012) Quantification of the impact
of embedded graphite nanofibers on the transient thermal response of paraffin phase change
material exposed to high heat fluxes. J Heat Transf 134:071901
6. Cai Y, Gao C, Zhang T, Zhang Z, Wei Q, Du J, Hu Y, Song L (2013) Influences of expanded
graphite on structural morphology and thermal performance of composite phase change
materials consisting of fatty acid eutectics and electrospun PA6 nanofibrous mats. Renew
Energy 109:163–170
7. Sari A, Bicer A (2012) Thermal energy storage properties and thermal reliability of some
fatty acid esters/building materials composites as novel form stable PCMs. Sol Energy Mater
Sol Cells 101:114–122
References 47

8. Grodzka PG (1970) Study of phase-change materials for a thermal control system. Lockheed
Missiles & Space Company, Denver
9. Warzoha R, Fleischer AS (2014) Improved heat recovery from paraffin-based phase change
materials due to the presence of percolating graphene networks. Int J Heat Mass Transf
79:314–323
10. Javani N, Dincer I, Naterer GF, Yilbas BS (2014) Heat transfer and thermal management with
PCMs in a Li-ion battery cell for electric vehicles. Int J Heat Mass Transf 72:690–703
11. Zalba B, Marin JM, Caberza LF, Mehling H (2003) Review on thermal energy storage
with phase change: materials, heat transfer analysis and applications. Appl Therm Eng
23:251–283
12. Kenisarin M, Mahkamov M (2007) Solar energy storage using PCMs. Renew Sustain Energy
Rev 11:1913–1965
13. Sharma SD, Sagara K (2005) Latent heat storage materials and systems: a review. Int J Green
Energy 2:1–56
14. Denholm P, Mehos M (2011) Enabling greater penetration of solar power via the use of CSP
with thermal energy storage. NREL technical report:NREL/TP-6A20-52978
15. Robak CW, Bergman TL, Faghri A (2011) Economic evaluation of latent heat thermal energy
storage using embedded thermosyphons for concentrating solar power applications. Sol
Energy 85:2461–2473
16. Feldhoff JF, Schmitz K, Eck M, Schnatbaum-Laumann L, Laing D, Ortiz-Vives F, Schulte-
Fischedick J (2012) Comparative system analysis of direct steam generation and synthetic oil
parabolic trough power plants with integrated thermal storage. Sol Energy 86:520–530
17. Rathod MK, Banerjee J (2013) Thermal stability of phase change materials used in latent
heat energy storage systems :a review. Renew Sustain Energy Rev 18:246–258
18. Peng Q, Wei X, Ding J, Yang J, Yang X (2008) High temperature thermal stability of molten
salt materials. Int J Energy Res 32:1164–1174
19. Kimura H, Kai J (1984) Phase change stability of CaCl2·6H2O. Sol Energy 33:49–55
20. Ge H, Liu J (2013) Keeping smart phones cool with gallium. J Heat Transf 135:054503
21. Ge H, Li H, Mei S, Liu J (2013) Low melting point liquid metals as a new class of phase
change material: an emerging energy frontier. Renew Sustain Energy Rev 21:331–346
22. Khare S, Dell’Amico M, Knight C, McGarry S (2012) Selection of materials for high tem-
perature latent heat energy storage. Sol Energy Mater Sol Cells 107:20–27
23. Kensarian MM (2010) High-temperature phase change materials for thermal energy storage.
Renew Sustain Energy Rev 14:955–970
Chapter 4
PCM Design Issues

4.1 Overview of PCM Implementation

We have seen that PCMs come in many different types, from inorganic salts to
liquid metals, and can be applied in applications as diverse as spacecraft design
to domestic hot water tanks. However, in implementation, all of the PCM types
and applications share certain attributes and design challenges. Of course, the pri-
mary characteristic that all PCMs share is that they transition from solid to liquid
over the extent of the operating time. Without this transition the latent heat would
not be exploited, and the energy storage or thermal control would rely solely on
sensible heating. However, the transition to a liquid state creates specific design
challenges focused on the containment of the liquid phase in such a way that the
system reliability and performance are not affected. In most cases the PCM will
have to be located within some sort of packaging, and this chapter discusses sev-
eral of the most popular containment methods.
The transition to liquid phase in most cases also results in a significant den-
sity change. Though certainly not of the scale seen during transition from liquid
to vapor, the solid-liquid phase change process results in a volume change of up
to 15–20 % for most common PCMs. This volume change must be considered and
accounted for when designing PCM systems.
Finally, in most PCM applications, the response time is a significant challenge.
The PCM needs to be located directly in the heat flow path to promote quick pen-
etration and initiation of melting. However, the low thermal conductivity of most
popular PCMs, as noted in this chapter, leads to a high thermal resistance which
diverts the heat flow into alternate lower resistance paths, creating heat flow bypass
of the PCM. Thus in application, most designs utilize some form of thermal con-
ductivity enhancement. The first part of this chapter discusses several of the most
common methods of conductivity enhancements and their advantages and disad-
vantages, while the second half of the chapter details various containment designs.

© The Author(s) 2015 49


A.S. Fleischer, Thermal Energy Storage Using Phase Change Materials,
SpringerBriefs in Thermal Engineering and Applied Science,
DOI 10.1007/978-3-319-20922-7_4
50 4  PCM Design Issues

4.2 Enhancement of Thermal Conductivity

The enhancement of the thermal conductivity of phase change materials is one of


the most pressing topics in PCM design and implementation. It is widely accepted
that the low thermal conductivity of materials with high energy storage capabili-
ties is one of the most significant barriers to the wider implementation of thermal
energy storage applications. A significant amount of development work is focused
on this topic, with the goal of creating high latent heat, high specific heat, high
thermal conductivity materials.
This is a critical path to implementation because the use of energy storage
materials with low thermal conductivities can lead to degraded system perfor-
mance. For example, when used for the thermal control of portable electronic sys-
tems, the response time of the material to absorb the thermal pulse is important.
If the thermal resistance is such that the heat flux cannot quickly diffuse into the
PCM to initiate melting, instead of acting as a thermal sink the PCM may instead
act as a thermal insulator. With thermal conductivities as low as 0.2 W/m K, the
popular family of paraffins can easily act as a thermal insulating layer if steps are
not taken to improve thermal transport. In this case, when the processor is dissipat-
ing a lot of heat, the high thermal resistance prevents the heat from penetrating
into the mass of PCM and instead the melt front stalls right at the heat source.
The layer of PCM closest to the heat source melts, and the liquid layer quickly
superheats. This often induces natural convection, which will increase heat trans-
fer rates, but as the constant heat flux from the heat source continues to pump into
the thin liquid layer, the source temperature continues to rise, leading to unsafe
operating conditions, as seen in Fig. 4.1. To avoid this scenario, it is necessary to
provide low resistance heat flow paths that promote effective penetration of heat
into the wider mass of PCM, avoiding superheat of the liquid layer right at the
source. An additional challenge is to provide these heat flow paths and/or conduc-
tivity enhancements without impeding natural convection, which increases the
melting rate and is advantageous to the application in most cases.
In solidification, which has been shown to be the slower, conduction domi-
nated process, the low thermal conductivity of the PCM can significantly impede
the heat flow. In this case, the PCM solidifies on the cold surface, meaning that
the heat from the phase change process must penetrate an increasingly thick solid
layer as seen in Fig. 4.2. In this case, the thermal resistance grows ever higher as
the heat flow path from the hot liquid phase through the subcooled solid phase
increases with time. It is extremely important to provide enhancements that
increase the solid layer conductivity in order to avoid excessively long solidifica-
tion times that preclude the use of these materials in fast acting transients.
There are three popular methods to increase the effective thermal conductivity
of PCM. The first is the use of macroscale metallic inclusions such as fins, meshes
or foams. The second is the use of macroscale carbon inclusions, which utilizes
the high intrinsic thermal conductivity of carbon based materials, and the final
method is the use of nanoscale materials to create colloidal PCM suspensions with
improved thermal properties.
4.2  Enhancement of Thermal Conductivity 51

Fig. 4.1  Heat pumping into a superheated liquid layer

Fig. 4.2  Solidification on a cold surface with an increasingly thick solid layer that impedes heat
transfer
52 4  PCM Design Issues

4.2.1 Metallic Inclusions

The high thermal conductivity of most metals makes the use of metallic inclusions
to improve the thermal response time of PCMs an obvious design choice. In these
systems, variously shaped metal parts can be embedded within the PCM to pro-
vide heat flow paths through the entire PCM mass, creating a higher effective ther-
mal conductivity of the system.
One of the simplest designs is to embed a finned heat sink within the PCM con-
tainer. The fins enable the heat to flow rapidly along the high conductivity spines
enabling heat penetration deep within the container. The heat then dissipates off
the fins into the PCM as seen in Fig. 4.3. The fins effectively separate the PCM
mass into smaller, thinner layers leading to easier diffusion of the heat through the
PCM. The extra surface area that the fins provide is highly effective at removing
heat rapidly from the heated base of the system, and the interaction of conduction
through the fin spine and natural convection in the melt enhances the overall sys-
tem performance. Orientation of the system such that natural convection will initi-
ate upon the initiation of the melt process is preferred if possible.
Embedded heat sinks have been shown to be highly effective in PCM systems,
and the size of the system does not necessarily have to be large. For instance, a
numerical parametric study showed the impact of a plate fin heat sink with fins
5–10 mm high and 0.15–1.2 mm thick [1]. In this case, the PCM filled the space
between the plate fins which was varied parametrically from 0.5–4 mm thick. The
heat sink base was horizontal (as in Fig. 4.3) and the base temperature varied from
6–24 °C above the melting point of the PCM. The model clearly showed that the

Fig. 4.3  Heat sink embedded


in PCM with heat flow path
shown
4.2  Enhancement of Thermal Conductivity 53

heat penetrated quickly along the length of the fin and then into the PCM, with
a much lower proportion of the heat moving directly from the baseplate into the
PCM. However, as the space between fins increased, and the PCM layer thickness
increased correspondingly, then the proportion of heat dissipated directly from the
baseplate became larger.
The motion of the melt front in this case also showed the effectiveness of the
fins in promoting rapid thermal diffusion [1]. For widely spaced fins, the melt
front moved away from each fin face in a parallel direction towards the next fin as
shown in Fig. 4.4a. However, as the fin faces were spaced more closely together,
the melt front motion began to show an effect of the baseplate-induced melt-
ing leaving a solid wedge of PCM closer to the top of the enclosure as seen in
Fig. 4.4b.
The larger the temperature difference between the baseplate and the PCM melt
point, the faster the PCM melted. Similar behavior is seen for constant heat flux
conditions [2]. The aspect ratio of the plate fins is also found to have a strong
effect on the system performance [3]. The aspect ratio here is defined as the height
of the fins divided by the gap between fins, so a large aspect ratio reflects tall
tightly packed fins, while a small aspect ratio features short, widely spaced fins.
The low aspect ratio designs were found to dissipate most of their heat through
the baseplate with the melt front almost parallel to the baseplate, while the high
aspect ratio design were found to be conduction dominated due to the extremely
thin PCM layers featured in these designs.
These models all showed the effectiveness of embedded heat sinks at pro-
moting heat diffusion into PCM containers, and similar results have been seen
in experimental work [4]. However, embedded heat sinks do have several disad-
vantages including added weight, and more importantly, the displacement of a

Fig. 4.4  Melt front motion between fins a parallel for widely spaced fins, b impacted by base-
plate melting for tightly spaced fins
54 4  PCM Design Issues

significant volume of PCM from the container, which reduces the energy storage
capacity. For these reasons, other methods of enhancing conductivity with lower
volume/mass options have been explored.
Metallic foams, which feature large surface area/mass ratios have also been
shown to be effective at improving the effective thermal conductivity and diffusiv-
ity of PCMs. The use of metallic foams is similar in nature to the use of embedded
heat sinks, in that the heat flows along a metallic path, in this case the ligaments to
the foam, and then into smaller separated masses of PCM. For foams, the individ-
ual masses of PCM are that which are contained within each open cell in the foam.
This can be seen in Fig. 4.5.
Various metallic foams, most commonly aluminum [5], although also copper
[6] and nickel foams [7], have been studied for their effectiveness in improving
the thermal response of PCMs. The use of aluminum foam is shown to signifi-
cantly reduce temperature variations within the PCM mass and to develop a more
uniform melt process due to the effective heat spreading of along the ligaments
[5]. In a case using an aluminum foam sample featuring 40 pores per inch, the
melt process was seen to be conduction dominated [5], similar to that seen with
tightly packed fins. The suppression of natural convection in the melt can possibly
lead to increased temperatures at the heated base during melting. Experimental
studies have shown though, that while the convection may be suppressed, the
overall thermal performance is still significantly better with foams than without.
In a study using paraffin PCM and copper foam of three different designs, rang-
ing from 10–30 PPI, data showed that while convection was suppressed, the tem-
perature differences within the PCM were reduced, the base temperature control
was better and the melt was much more uniform than for the case without any
foams [6]. During the solidification process, which is conduction dominated as
explained in Sect. 4.2, the use of metallic foam decreases the solidification time
of the PCM.

Fig. 4.5  Porous aluminum
foam filled with paraffin
PCM
4.2  Enhancement of Thermal Conductivity 55

Numerical studies have been able to accurately capture the physics of the inter-
action between the metallic foams and the embedded PCM [8, 9]. These models
have confirmed that the heat spreads rapidly through the foam and then diffuses
more slowly into the PCM in each pore through a conduction dominated process
[8]. However, if the foam is porous enough and the temperature difference high
enough, convection may occur even within the foam, leading to enhanced heat
transfer in these situations [9].
The effect of embedded foams on system thermal conductivity has been directly
measured [7]. In this study, copper and nickel foams with porosities between 88
and 96 % and pores per inch from 5–25 were filled with paraffin PCM with a
melt temperature of 62 °C under vacuum conditions to ensure a complete fill. The
results showed that the effective thermal conductivity of the paraffin/foam compos-
ites increased as the porosity decreased and as the pores per inch increased, in both
cases due to a higher metal/PCM ratio. The copper foams showed greater effec-
tive thermal conductivities than the nickel foams, reaching as high as 16 W/m K
for copper foams with 25 pores per inch and 88 % porosity [7]. Considering that
the thermal conductivity of paraffin is ~0.2 W/m K, this improvement is almost
80×. However, the lower porosity and higher pores per inch, leads to a lower mass
of paraffin in an equivalent volume, reducing the energy storage capacity. The
reduced energy storage capacity must be considered by the designer when deciding
whether it is a worthwhile trade-off for the improved thermal diffusion.

4.2.2 Macroscale Carbon Inclusions

Although the previous section clearly illustrated the effectiveness of metallic


inclusion such as fins and foams on the thermal response of PCMs, the high ther-
mal conductivity of most carbon based materials also makes them a popular and
effective choice for PCM enhancement. There are various types of macroscale car-
bon based materials that can be used, including carbon fibers and graphite foams.
Micro and nanoscale materials will be discussed in the next section.
Carbon fibers have been shown to be highly effective at increasing the effective
thermal conductivity of PCMs [10]. Carbon fibers in general have high thermal
conductivities along the length of the fiber and are resistant to corrosion or chemi-
cal attack, making them effective for long term use. In a well-known experimental
study [10], carbon fibers with a diameter of 10 mm, a density of 2170 kg/m3 and
a thermal conductivity of 220 W/(m K) were embedded in two different configu-
rations in a paraffin PCM with a melt temperature of 41 °C and a base thermal
conductivity of 0.26 W/m K. The composite material was packed in a cylindrical
container 50 mm in diameter and 130 mm high. In the first case, carbon fibers with
lengths between 5 and 200 mm were randomly dispersed with the PCM. The dis-
tribution in fiber length resulted in many fibers being longer than either the diam-
eter and/or the capsule height, leading to coiling or tangling of the fibers with the
container. The second design featured a “brush type” design with the carbon fibers
56 4  PCM Design Issues

Fig. 4.6  Randomly
dispersed carbon fibers and
brush type carbon fibers

attached to an axis which was located at the center of the capsule. The carbon fib-
ers then extended radially from the axis to the edge of the container with all fibers
exactly 25 mm in length as seen in Fig. 4.6.
The volume fraction of fibers was set to 2 % because it was found that during
the melt process, even low volume fractions of fibers suppress the development
of natural convection in the melt. This initially decreases the overall heat transfer
during melting, but as the volume fraction rises to 2 %, the increase in conduction
heat transfer more than offsets the reduction in convection heat transfer. This is
similar to what was seen in the metal foams, where the system was more effective,
even with the natural convection suppression.
During the solidification process, the addition of fibers decreases the solidifica-
tion time and shows a benefit for all fiber loading ratios as convection does not
play a significant role during solidification. In the solid phase, the 2 vol% addi-
tion of carbon fibers increased the effective thermal conductivity of the composite
by 6 times for the randomly dispersed samples. Little to no effect of fiber length
was seen on the thermal conductivity or the thermal performance. However, the
effect of fiber orientation was found to play a significant role. The capsule was
cooled from the outside to the center, so the brush-type fibers provided a direct
conduction path in the direction of heat flow. This lead to an increase in thermal
performance featuring faster melting and solidification, and an increase in effec-
tive thermal conductivity of 3× over the randomly orientated fibers.
Graphite foams are also a popular choice to enhance the thermal response of
PCMs. As with metallic foams, graphitic foams feature high surface area to vol-
ume ratios, leading to enhanced heat transfer performance. However, unlike the
uniform, easily controlled pores of most metallic foams, as seen in Fig. 4.5, many
graphite foams instead have irregular pores with ragged edges as seen in Fig. 4.7
4.2  Enhancement of Thermal Conductivity 57

Fig. 4.7  SEM images of graphite foam pores

and may even feature a number of closed cells within the foam [11, 12]. This can
make it difficult to effectively fill graphite foams with PCM. As such, one must
control the impregnation process much more carefully with graphite foam than
with metallic foams.
Metallic foams are often filled with PCM by simply pouring liquid PCM into
the foam within a container, or by submersing the foam within a liquid PCM
bath and letting the PCM infiltrate the pores. However, these methods are inef-
fective with graphite foams, and lead to incomplete fill of the foam. A study of
effective methods for infiltration of PCM into graphite foams [11] showed that the
submersion technique led to fill ratios of 53–81 % when the PCM is in the liquid
phase, meaning that 19–47 % of the open space in the foam remains devoid of
PCM. A better method of infiltration is to use a vacuum oven to remove all the air
within the foam prior to melting the PCM and allowing it to wick into the evacu-
ated foam. This method has been shown to achieve fill ratios of 97–100 % [11].
Another study [12] showed that the microcracking characteristic of some graphite
foams also helps to improve the infiltration rate as many foams feature few of the
large pores which would make infiltration easier.
When properly filled, graphite foam has been shown to be highly effective at
improving thermal conductivity and diffusivity in PCMs. A study on the use of
graphite foam filled PCM for thermal management of electronics [5] showed that
graphite foam was more effective than either aluminum foam or graphite fibers at
controlling the heated base temperature over the time of the transient temperature
pulse and the temperature distribution within the PCM was nearly uniform due
to the high effective thermal conductivity. A study on the effect of porosity and
the ligament structure of the graphite foams [13] showed the thermal diffusivity
of the foam/PCM composites increases as the ligaments become thicker and the
pore size decreases. However, as for the metal foams discussed earlier, increasing
the ligament thickness and decreasing the pore size leads to lower masses of PCM
and decreased energy storage ratios. This is shown quantitatively by Song et al.
[12] who showed that the latent heat of graphite foam/paraffin composites dropped
from 110 to 65 kJ/kg as the foam density increased from 0.2 to 0.57 g/cm3.
58 4  PCM Design Issues

The use of graphene aerogels, which are among the lightest materials in exist-
ence, is being explored as a possibility to create highly effective graphite/PCM
composites without displacing a significant amount of PCM from the compos-
ite and reducing the overall latent heat/energy storage capability. Single layer
graphene features one of the highest known thermal conductivities (4800–
5300 W/m K) [14] and few layer graphene particles also exhibit extremely high
thermal conductivities. In their simplest form, graphene aerogels may be thought
of as a three-dimensional form of few layer graphene, with the graphene sheets
forming the structure of a macroscale foamed structure [15].
The use of aerogel as a supporting matrix for PCM is gaining traction and a
recent paper by Zhong et al. [15] explores the performance of a graphite aerogel
impregnated with octadecanoic acid PCM. The aerogel was fabricated in-house
for this study and features a density of 0.227 g/cm3 and a thermal diffusivity of
8.580 mm2/s. The thermal diffusivity of the octadecanoic acid is 0.095 mm2/s. The
composite PCM/graphene aerogel material was found to have a thermal conductiv-
ity 13× greater than that of the octadecanoic acid alone, while the latent heat of
the composite dropped only slightly. The composite material had a latent heat of
181.8 kJ/kg while the octadeconoic acid has a latent heat of 186.1 kJ/kg, a reduc-
tion of only 2.3 %. Thus, while the research is preliminary, graphene aerogels
appear to be a promising avenue of work to develop high latent heat, high thermal
conductivity composite PCM materials.

4.2.3 Nanoscale Carbon Inclusions

Another promising avenue of research into the enhancement of thermal conductiv-


ity and thermal diffusivity of PCMs is the inclusion of micro and/or nanoscale car-
bon based materials within the PCM. These materials feature exceptionally high
intrinsic thermal conductivities, high surface area to volume ratios, and can exert
a significant impact on the thermal properties of the PCM/nanomaterial composite
at very low loading levels. The key to the successful use of these materials lies
in understanding how heat is transported not only within the nanomaterials them-
selves, but also between nanoparticles when used in a composite.
At the nanoscale, heat conduction within a carbon based nanoparticle is driven
by atomic lattice induced vibrations known as phonons. Phonons are energy pack-
ets, which in a simplified view can be imaged as lattice vibrations moving through
the material much in the same manner as a stretched out spring can transmit vibra-
tions from one end to another. The more tightly bound the crystalline structure, the
more easily the phonons traverse the length, transporting energy from one location
to another through nanoscale conduction. Phonons can be scattered and dispersed
through interactions with defects in the material such as vacancies in the crystal
structure, at the boundaries of the materials, and through interaction with other
phonons.
4.2  Enhancement of Thermal Conductivity 59

Carbon allotropes exhibit extremely tight covalent sp3 or sp2 bonds between
individual atoms which allow phonons to move rapidly through these materials
with little attenuation. This leads to extremely high intrinsic thermal conductivi-
ties. The exact value of thermal conductivity for different carbon based structures
such as single-walled carbon nanotubes, (SWCNT), multi-walled carbon nano-
tubes (MWCNT), graphite nanofibers (GNF) and graphene nanoplatelets (GNP)
is a matter of some debate [16], but it is widely recognized that these materials
exhibit exception thermal properties, with thermal conductivities cited from 300–
3000 W/m K depending on the material. Examples of these materials can be seen
in Fig. 4.8. Single layer graphene (SLG) is formed as its name suggests from a
single atomic layer of high quality graphene and is the 2-D structure on which
all the other forms are based. Graphene nanoplatelets or few layer graphene are
formed from several layers of stacked SLG creating a particle 1–15 nm in thick-
ness. Single and multiple walled CNTs are formed from a tubular version of gra-
phene sheets, while GNF feature many layers of graphene arranged along an axis
into a long fiber which can be micrometers in length.
The intrinsic thermal conductivity is affected by the mean free path of the pho-
nons within the nanoparticle. Long mean free paths between phonon scattering
events lead to higher thermal conductivities. For instance in SWCNT, if the mean
free path of the phonon is longer than the length of the SWCNT, then the transport
regime is known as ballistic and there is little scattering along its length leading
to high thermal conductivities. If however, the mean free path is shorter than the
length, the regime is known as diffusive, and the increased scattering events lead
to lower thermal conductivities. In general, the more opportunity for scattering to
occur at internal defects, the lower the thermal conductivity, thus intrinsic thermal
conductivity decreases as diameter and length increase. Scattering also increases
with more boundaries, whether within the particle, such as between graphene

Fig. 4.8  Carbon based nanoparticles from left to right: SWCNT, MWCNT, SLG, GNP
60 4  PCM Design Issues

sheets in MWCNT and GNF, or between particles. Thus, the thermal conductiv-
ity of PCM/nanoparticle composites should be highest when using nanoparticles
with little internal scattering and when the interfaces between nanoparticles are
minimized.
The thermal properties of PCMs embedded with SWCNTs, MWCNTs and
GNF have been quantified by several research teams [17–20]. In all cases, the
presence of the nanoparticles was found to improve the thermal conductivity of the
material to varying degrees. The effect of multiwalled CNTs was studied by Elgay
and Lafdi [17]. This study examined the impact of embedded MWCNTs with an
average diameter of 100 nm and an average length of 20 μm into paraffin PCM
with a melt temperature of 67 °C at weight percentages (wt%) from 1–4 wt%.
The increase in thermal conductivity was more than 35 %, from just below
0.25 W/m K for the base paraffin to about 0.33 W/m K at 4 wt%. The increase in
conductivity was linear with loading level.
A similar increase in thermal conductivity was seen for 30 nm diameter, 50 μm
long MWCNT embedded in paraffin with a melt temperature of 53 °C [18]. The
paraffin/MWCNT composites had loading levels of 0.2–2 wt%, and the highest
thermal conductivity was recorded for the highest loading level. At 2 wt%, the
composite showed a 35 % improvement in thermal conductivity to 0.32 W/m K.
In this case it was also shown that the addition of the MWCNT slightly decreased
the latent heat and the melting temperature. However, this study examined only
the thermal properties of the composite material and did not look at the thermal
response of the materials in operation.
Several studies have looked at thermal cycling behavior of PCM/nanoparticle
composites [19, 20]. These studies examined the impact of PCMs with embedded
graphite nanofibers on the melting cycle [19] and the solidification cycle [20]. The
PCM/GNF composites were made using herringbone style fibers in which the par-
allel planes of graphene are shaped like a peaked roof rather than a flat plate. The
GNF had diameters up to 100 nm and lengths up to 100 μm and were embedded
in the PCM at 5–11 wt%.
During the heating cycle, the high loading levels of GNF were found to reduce
the temperature spread within the composite material as it heats up to the melt
temperature due to its higher thermal conductivity [19]. However, once the melt-
ing point was reached, the high loading levels suppressed the natural convection
within the PCM, much as the low porosity foams did (see Sect. 4.2.2). This led
to overheating at the base, despite the higher thermal conductivity and diffusivity,
thus these materials need to be optimized not only for their thermal properties, but
also their thermal performance as a system during cycling conditions.
During solidification [20], the thermal performance was improved, due to the
conduction dominated nature of the solidification process. In fact, a loading level
of 10 wt% GNF led to a 61 % reduction in total solidification time. For systems
which need to feature a quick “recharge” of the PCM for the next transient pulse,
this can be highly advantageous.
A comparison of composites made from four different carbon based nano-
structures showed that graphite nanoplatelets lead to the overall highest thermal
4.2  Enhancement of Thermal Conductivity 61

conductivity improvements [21]. In this study, a paraffin PCM was embedded with
short MWCNT, longer MWCNT, GNF and GNP at loadings up to 5 wt%. The
shorter MWCNT based composites outperformed the GNF and longer MWCNT
composites, mainly due to the greater phonon attenuation in the longer MWCNTs
and GNFs. The 5 wt% SWCNT composites exhibited a thermal conductivity
increase from 0.263 to 0.324 W/m K, a 23 % increase. The longer MWCNT com-
posite exhibited an increase of only 17 % and the GNF composite 16 %. More
interestingly though, the 5 wt% composite based on GNP had an increase in
thermal conductivity of 166 % to 0.7 W/m K, significantly higher than the other
materials. The 2-D GNP thus exhibited a much stronger influence on the overall
composite thermal conductivity even at similar loading levels.
A similar effect was found in another study [22] which examined exfoliated
graphite nanoplatelets (xGNP) at loading levels of 1–7 wt% in paraffin with a
melt temperature range of 53–57 °C. At 5 wt% the thermal conductivity was again
found to be 0.7 W/m K, which then increased to 0.8 W/m K at 7 wt%.
It appears that in order to optimize the thermal conductivity of the material, it is
best to use carbon based nanoparticles that are pancake shaped like the GNP rather
than tube shaped like SWCNT due to the decreased number of interfaces between
particles when working with GNP. Interfaces with a larger surface are per interface
as seen in Fig. 4.9 leads to lower phonon scattering between particles which leads
to increased heat transport. In fact, interfaces between SWCNT can constrict the
phonon transport to perhaps only a few atoms in contact.
The effect of interfaces on the thermal behavior of PCM composites was shown
clearly in a comprehensive study which examined the thermal properties of PCM
composites made with 9 different types of xGNP [23]. The xGNP varied in both
thickness (2, 6 and 15 nm) and width/diameter (1.5, 5, 15, and 25 μm). The load-
ing levels in the PCM varied from 0.01–20 vol%, which spanned the range across
which the particles when dispersed within the PCM would not be in intimate con-
tact with each other (dilute) to when they would be in constant contact with each
other across the composite (percolating).

Fig. 4.9  Interface geometry between CNTS and GNPs


62 4  PCM Design Issues

The results showed significant increases in thermal conductivity for all com-
posites. In the dilute regime (<1 vol%), the increase was modest for the thinnest,
smallest particles which have the largest number of particle-particle contacts.
These particles showed an increase in thermal conductivity of up to 20 %. The
thicker, wider GNP had larger increase in thermal conductivity of up to 80 % in
this same region [23].
At higher loading levels, the particles begin to percolate, providing a continu-
ous path across the composite from particle to particle. In this regime, the smallest
thinnest particles once again showed the smallest increase in thermal conductivity.
With the particles in continuous contact the thermal conductivity rose to 1 W/m K,
almost a 400 % improvement from the base PCM (0.26 W/m K), but this increase
is slight compared to what was seen with the larger particles. In fact, for the
thickest, most rigid GNP with thickness of 15 nm and width/diameters of 25 μm
showed an increase in thermal conductivity in the percolating region to almost
7 W/m K, an improvement of 25×.
Based on phonon physics, the intrinsic thermal conductivity should be highest
for the thinnest smallest GNP, leading to the highest composite thermal conductiv-
ity. As this is not the case, it appears that the number of contact points between
nanoparticles plays a dominant role in the overall thermal conductivity of the com-
posite material.
Based on this overview of carbon nanoparticle based PCM composites, it is
clear that while SWNCT, MWCNT and GNF can offer improved thermal proper-
ties, the best performance seen to date occurs with composites made with GNP
due to their exceptional phonon transport, and lower number of interface con-
tacts when used in a matrix. Greater improvement is seen when the materials are
allowed to percolate, although the benefit will be highest in solidification as the
percolating nanostructures will suppress convection in the liquid phase. The GNP
based materials have not yet been tested in thermal cycling so the impact during
the melt phase is not yet clear. These materials appear to offer significant benefits,
but further development work is necessary before implementation.

4.3 PCM Containment Issues

4.3.1 Basic Designs

There are several different ways to implement PCM in a given application such
that the liquid melt phase remains contained and does not contaminate the sys-
tem. This chapter will discuss the design and use of macroscale container systems,
of microencapsulated PCM beads, and of completely form stabilized PCM. All of
these systems are similar in that their goal is to completely contain the liquid PCM
within a capsule or shell of some sort, made of a material which does not change
phase. This shell can be as simple as an aluminum box, or as complex as a chemi-
cally synthesized polymeric composite nanosphere.
4.3  PCM Containment Issues 63

Regardless of complexity, one attribute that all the containers share is that they
must deal with the issue of PCM volume change during the melting process. Most
PCMs exhibit a volume change of up to 15–20 % upon phase change with the solid
phase typically denser, and thus lower in volume than the liquid phase. This results
in an empty space within the container while the material is in its solid phase. This
empty phase is referred to as an ullage space. The ullage space must be considered
during the design of whichever container style is appropriate for the application.
The ullage space will create an air gap in the container which, depending on its
location, can result in a degradation of thermal performance. An example of ullage
space formation in an aluminum box containing paraffin can be seen in Fig. 4.10.
The thermal conductivity of air is several orders of magnitude lower than that of
any PCM and creates a high thermal resistance that can divert the heat flow to
alternate paths. Therefore the casing must be designed to promote the formation of
the ullage space in a location away from the heat source. It has been shown that the
location and size of the ullage space can significantly affect the temperature gradi-
ents and melt ratios within a PCM container, impeding heat flow and reducing the
effectiveness of the energy storage capabilities [24]. This means that under normal
operating conditions where gravity acts as a downward force, (i.e. not space based
applications) that situations with the heat source on the top plane of the PCM con-
tainer should be avoided, as the ullage space will form at the top of the container
and the heat source will be insulated from the PCM by the air layer.

Fig. 4.10  Solid paraffin
exhibiting an ullage space
in an aluminum container
64 4  PCM Design Issues

4.3.2 Macro-encapsulated PCMs

The simplest way to enclose PCM so that the liquid phase remains contained is to
use a rectangular box, fabricated from a material that is compatible with the cho-
sen PCM, and is high conductivity to reduce thermal resistance. The box should be
filled with liquid PCM to the top of the container, the lid replaced, and the PCM
allowed to cool. The compatibility of the PCM and the container material must be
kept in mind at all times as some PCMs are corrosive to certain common materi-
als. For instance, hydrated salts have a number of compatibility issues, and liq-
uid metals such as gallium are often corrosive. In a series of tests, stainless steel
exhibited corrosion resistance to the greatest number of PCMs [25], although its
low thermal conductivity compared to other materials is a deterrent.
Under repeated heating and cooling cycles the liquid PCM will expand and con-
tract while remaining sealed within the box. The low viscosity and/or high melt
temperature of certain PCMs can make sealing a challenge, so the thermal engineer
will have to carefully design the box to avoid leakage. High temperature gaskets
and sealants can be effective, but in other cases leakage can become a significant
problem, and the containers should be permanently sealed by welding, brazing and/
or soldering if appropriate for the material. Containers should be fabricated with as
few leak points as possible for this reason. In Fig. 4.10, thermocouples on the right
hand side are sealed with high temperature sealant to avoid leakage. Chintakrinda
et al. [5] uses aluminum boxes with different types of embedded matrices such as
aluminum and carbon foams to improve thermal conductivity. Several containers
filed with carbon foam can be seen in Fig. 4.11. Upon close examination it can be
seen that some of the containers exhibit welding seams along vertical joints, while
others are formed from aluminum channel to minimize leak points.

Fig. 4.11  Aluminum
containers with carbon foam
4.3  PCM Containment Issues 65

Cylindrical or spherical containers can be used instead of rectangular or cubic con-


tainers to minimize leak points. For instance, large cylindrical capsules were used in
a concentrating solar application to contain sodium nitrate (NaNO3) and a eutectic
mixture of sodium chloride (57 %) and magnesium chloride (43 %) [26]. These mate-
rials have high melting points (308 and 444 °C respectively) and can have compat-
ibility issues with many possible container materials. Testing showed that these salt
mixtures could be contained using 304 stainless steel without excessive corrosion,
but testing was not conducted for long-term durability. Two container designs were
used—38.1 mm diameter by 101.6 mm tall, and 25.3 mm diameter by 46.4 mm tall.
An ullage space of 18–20 % volume was found to form at the top of each capsule.
In situations where structural strength is not an issue, thin film plastic bags can be
used. The use of flexible thin film packaging reduces issues with ullage space con-
traction as the thin film is flexible enough to accommodate the volume change For
instance, polyethylene thin film packaging was used to contain paraffin for energy
storage in an under-floor ventilation duct system with the intention of reducing build-
ing energy consumption [27]. In this work multiple “bags” of PCM were supported on
a wire rack. The rack provided structural support for the flexible containers while also
easily allowing air flow over and around the PCM. Potential concerns with this system
include the risk of rupture of the thin film bags, abrasion, and aging of the plastic,
reducing its flexibility. Long term testing of this type of containment is needed.
Containment in PCM packed bed designs typically uses hard shell spheres.
Spheres have the advantage of a large surface area to volume ratio, and the pack-
ing ratio of spheres in a tank promotes air flow between and around the spheres,
promoting high heat transfer rates. The spheres are usually quite large. For
instance, a packed bed used in a thermal storage tank for a solar heating system
prototype [28] used 75 mm diameter spheres molded from high density polyeth-
ylene filled with a commercially available PCM referred to as HS58 with a melt/
freeze range of 55–58 °C. The storage tank held 500 carefully packed spheres
with 250 g of PCM each for a total energy storage capacity of 31,250 kJ. The tank
was charged using a solar air heater as seen in Fig. 4.12. At full charge (all PCM

Fig. 4.12  PCM containment
in spheres for solar air
heating packed bed design
66 4  PCM Design Issues

fully melted) the tank was shown to be capable of heating ambient air to 55 °C for
2–4 h depending on air mass flowrate.
All of the macroscale containment methods discussed here can be combined
with the various thermal conductivity enhancement methods presented earlier in
this chapter, for flexibility to adapt the design to the specific application.

4.3.3 Micro-encapsulated PCMs

Microencapsulation of PCMs avoids the bulkiness of the large macro scale cap-
sules, and instead features microscale beads with polymeric shells which encapsu-
late a PCM core as seen in Fig. 4.13. These beads are referred to as mononuclear
microcapsules [29]. The scale of the material is typically 1–1000 μm in diameter
for microPCM beads and 10–500 nm for nanoPCM beads. While these beads hold
a significantly smaller mass of PCM than the macroscale containers discussed in
the previous section, the small size offers some distinct benefits. One of the pri-
mary benefits is the ability to easily embed these beads in other materials for ease
of fabrication of energy storage materials. As discussed in Chap. 2, microencap-
sulation beads are used to fabricate wallboard, concrete, roofing materials and
fabrics with energy storage capabilities, by either mixing the beads into the mate-
rial as it is being fabricated, or by spraying the beads on as a post processing step
using an adhesive coating. The beads can even be entrained in liquids to create
heat transfer fluids with additional energy storage capability. Another benefit is
their high volume to surface area ratio which leads to a large heat transfer area and
a fast thermal response with shortened melt and solidification times.

Fig. 4.13  Microencapsulated
PCM
4.3  PCM Containment Issues 67

There are several different ways to create microencapsulated PCM using both
physical and chemical methods, although chemical methods are much more com-
mon. Possible physical methods can include air-suspension coating, the use of
vibrational nozzles and spray drying [29]. The chemical methods all involve differ-
ent forms of polymerization. Polymeric encapsulation is a common technique used
in many fields and is by no means limited to the development of micro-PCM. In
fact these techniques are commonly used in the food and pharmaceutical industries.
MicroPCMs can be formed using various well-known polymerization tech-
niques including interfacial polymerization, suspension polymerization and emul-
sion polymerization. All three techniques are based on the creation of a chemical
reaction that creates the rapid growth of a thin flexible polymeric coating on the
surface of a liquid bead of PCM. For instance, emulsion polymerization in a very
simplified manner can be thought of as occurring in three steps as seen in Fig. 4.14:
1. Small droplets of liquid PCM are dispersed in a carrier fluid.
2. A surfactant is added to the carrier fluid. The surfactant particles aggregate
and align into “micelles” such that the hydrophilic “heads” all face out and the
hydrophobic tails face in with the PCM particle in the center.
3. A reaction initiator is added to the carrier fluid where it reacts with the PCM
in the micelle, quickly polymerizing into a thin shell layer surround the PCM
particle.
Other variations of polymerization feature small differences in the basic proce-
dure. For instance, suspension polymerization requires continuous agitation dur-
ing the process. These techniques have been used quite successfully to encapsulate
different PCMs with various polymer shells.
Qiu et al. [30] used a modified suspension-based polymerization process to cre-
ate microencapsulated n-octadecane with crosslinked PMMA shells. The microbe-
ads were found to be over 75 % PCM. The PCM exhibited stable thermal behavior
and the encapsulation had little to no effect on its fundamental thermal behavior.
This study also pointed out the need to consider the strength of the shell in order
to prevent leakage during normal operation. The melting and solidification process
induces internal stresses on the shell due to the volume change, and the application
in which the material is used may exert external stresses as well. A high degree of
cross-linking during the shell formation process was found to create a shell with
higher mechanical integrity [30].

Fig. 4.14  Emulsion reaction creating microbeads of PCM


68 4  PCM Design Issues

With the shell integrity of upmost importance, it is important to understand that


the pH of the system in which the polymerization reaction occurs has a significant
effect on the formation of the microPCM beads. When the reaction occurs in an
acidic solution, the surface of the beads becomes very course and rough which can
lead to the formation of beads with an irregular surface profile, a higher possibil-
ity of damage and propensity for aggregation. Under basic conditions, the surfaces
instead tend to be smooth, and highly regular spheres are formed during the reac-
tion. This is due to the different manner in which the inorganic networks form in
acid and base conditions. A basic situation leads to the formation of dense and
highly cross-linked polymeric networks, with stronger shells [31].
Nanoscale PCM beads have also been successfully fabricated using similar
methods. NanoPCMs are of interest for use in heat transfer fluids, because the
addition of the larger microbeads of PCM can increase the fluid’s viscosity, and
are subject to significant shearing forces which damage the shell as they pass
through the pump or prime moved in the system. Nanosized PCM beads are less
likely to impact the viscosity or to sustain damage during pumping. Nanobeads
of n-octadecane with a PMMA shell were fabricated with an average diameter of
250 nm by Sari et al. [32] using an emulsion-based polymerization process. The
particles were found to be 43 wt% n-octacosane, a low encapsulation rate when
compared to microPCM.
NanoPCM was also successfully synthesized by Latibari et al. [33]. In this case
the PCM was palmitic acid with a shell of silicon dioxide using a sol-gel polymer-
ization process. The sol-gel process was also shown to be highly dependent on the
pH value of the solution, with the smallest diameters of 184 nm resulting when the
pH was maintained precisely at 11. As the pH crept upwards, so the bead diam-
eter rose quickly—to 466 nm at a pH of 11.5 and to 722 nm at a pH of 12. The
encapsulation ratios here were much higher than for Sari et al. [32], reaching 83 %
for even the smallest beads. Some of the smallest nanoencapsulated PCM beads to
date have been created by Zhang et al. [34] who fabricated nanospheres of octa-
decane with a PMMA shell of 119 nm in diameter using a mini-emulsion polym-
erization method and Kwon et al. [35] who fabricated nanocapsules as small as
63–84 nm using a mini-emulsion method with n-octadecane in PMMA.

4.3.4 Form Stable PCMs

An interesting way to contain PCM without the use of a container or shell is the
creation of shape-stabilized PCMs. Shape-stabilized PCMs are PCMs which retain
their shape even as the melt point is passed—direct observation of the material
would show no visible transition to liquid form, and no change in volume due to
the phase change. Of course, the PCM itself still transitions to a liquid as it moves
through the melt range, so what is happening with these shape-stabilized mate-
rials? SS-PCM, as it is called, is usually fabricated by mixing a liquid polymer
with liquid PCM. Both materials are heated well past their melt points and blended
4.3  PCM Containment Issues 69

together in specific combinations. The selected polymeric material should have


a melt point far above that of the PCM. High density polyethylene (HDPE) is a
popular choice when used with paraffin, as HDPE melts at 130 °C, well above the
55–60 °C for common paraffins. Upon through mixing the material is allowed to
cool, creating a polymeric PCM hybrid.
In application, the SS-PCM is used in systems where the operating tempera-
tures are above the melt point of the PCM, but below the melt point of the poly-
mer. As the polymeric PCM hybrid heats up above the melt point of the PCM,
the polymer remains solid while the PCM begins to melt. When the material is
properly designed, the polymeric material creates a solid scaffold which retains
the liquid PCM through capillary action within the cellular structure, and no melt-
ing is observed. Upon solidification, no ullage space is formed due to the poly-
meric support.
The advantage to the use of SS-PCM is the ability of these materials to melt
and solidify while retaining shape with no bulky containers. The materials can
be molded or machined into specific shapes, to allow easy adaptation to differ-
ent applications. Their disadvantage is the displacement of PCM in the material
in favor of the polymer. As the polymer does not change phase, the overall energy
storage of the material is reduced proportionally by addition of polymer. Thus it is
desirable to stabilize the PCM with as little polymer as possible.
Shape stabilization in paraffin-HDPE blends has been shown to occur above
25 % HDPE by Ehid and Fleischer [36]. In this work, paraffin-HDPE materi-
als were created in various combinations from 10 wt% HDPE to 50 wt% HDPE,
with 100 % paraffin and 100 % HDPE used as experimental controls. The materi-
als were tested to determine both thermal and structural behavior at temperature
below and above the melt point of the paraffin, and the material was considered
to be shape-stabilized when the material was found to retain its physical structure
with no leakage or seepage of paraffin from the structure at temperatures above the
melt point.
The hardness of the various paraffin/HDPE blends both below and above the
paraffin melt point can be seen in Figs. 4.15 and 4.16. Below the melt point, all
of the materials have some hardness and structural integrity, even the 100 % par-
affin sample, which is the softest of the samples. Addition of even 10 % HDPE
increased the hardness significantly (18 %). Further increases in HDPE resulted
in smaller increases in hardness, topping out at about a 25 % increase in hard-
ness. In comparison, at temperatures near or above the melt point of paraffin, in
this case operation at 50°, the 100 % paraffin and all blends below 25 % HDPE
lost their structure entirely and exhibited no hardness during impact testing. The
melting of the paraffin in this case was not contained by any HDPE structure.
However, above 25 % HDPE, it is clear that the structure was retained and the
material shape stabilized. Although the material softened considerably when com-
pared to the properties at ambient conditions, the material did retain its shape and
more importantly retained all the liquid paraffin without seepage or leakage.
The addition of the HDPE to a paraffin PCM also affects its thermal properties.
As noted earlier, the latent heat will be reduced proportionally to the percentage
70 4  PCM Design Issues

Fig. 4.15  Hardness of
paraffin/HDPE blends at
ambient conditions

Fig. 4.16  Hardness of
paraffin/HDPE blends above
the paraffin melt point

of HDPE in the final material as the polymer will not change phase. The specific
heat and thermal conductivities will also be affected. The specific heat will usually
be slightly reduced. For instance, the specific heat of HDPE can be as low as 1800
J/kg-K, while the specific heat of paraffin ranges from 2100 to 2900 J/kg-K. This
results in lower energy storage during sensible heating, as well as during phase
change. However, the thermal conductivity of the blended material will increase in
most cases. For example, the thermal conductivity of paraffin is about 0.2 W/m K,
while that of HDPE is more than doubled, ranging from 0.46 to 0.52 W/m K.
The addition of a stabilizer is advantageous when using nanoparticles of var-
ious types to increase conductivity further, as described in Sect. 4.2.3. The sta-
bilizing effect that the polymer exerts on the PCM also acts to stabilize the
nanoparticles, preventing any settling or agglomeration during repeated thermal
cycling. This was illustrated by Ehid et al. [37] using 70 % paraffin/30 % HDPE
blends with 5 wt% embedded graphite nanofibers (GNF). The 100 % paraffin
4.3  PCM Containment Issues 71

with GNF showed significant visible GNF settling after even one cycle, exhibited
changes in thermal response with repeated cycling. In contrast, the shape-stabi-
lized blend exhibited no visible GNF settling, and no change in thermal behavior
was evident even after 20 melting/solidification cycles.
The use of polymeric shape stabilization is not limited to paraffinic PCMs, and
an overview of different shape stabilized materials, including those based on par-
affin, fatty acids and polyethylene glycol is presented in a review by Kenisarin and
Kenisarina [38].
While polymeric shape stabilized PCMs are popular due to their effectiveness
and ease of ease of preparation, other types of SSPCMs are available. Any mate-
rial that can create a supporting structure which can retain PCM using capillary
forces can be used to create a SSPCM. Recently, embedded graphene oxide (GO)
sheets have been shown to not only enhance the thermal conductivity but also to
create a shape stabilized material when used in high enough percentages [39]. The
paraffin was impregrated into the GO sheets using a vacuum oven. The GO was
completely evacuated and then the PCM melted and wicked into the GO pores
over a 3 h period. Thermal testing of the material showed that a blend of 52 % gra-
phene oxide and 48 % paraffin was found to exhibit no liquid paraffin leakage over
2500 melting and solidification cycles, indicating shape stabilization. This load-
ing level of graphene oxide is somewhat higher than is seen for polymeric shape
stabilization, and the latent heat of the material is reduced accordingly. The use of
graphene oxide results in an increase in thermal conductivity, from 0.3 W/m K for
the selected paraffin to 0.985 W/m K for the paraffin/GO blend.

References

1. Shatikian V, Ziskind G, Letan R (2005) Numerical investigation of a PCM based heat sink
with internal fins. Int J Heat Mass Transf 48:3689–3706
2. Shatikian V, Ziskind G, Letan R (2008) Numerical investigation of a PCM-based heat sink
with internal fins: constant heat flux. Int J Heat Mass Transf 51:1488–1493
3. Saha SK, Dutta P (2010) Heat transfer correlations for PCM-based heat sinks with plate fins.
Appl Therm Eng 30:2485–2491
4. Hosseinizadeh SV, Tan FL, Moosania SM (2011) Experimental and numerical studies on per-
formance of PCM based heat sink with different configurations of internal fins. Appl Therm
Eng 31:3827–3838
5. Chintakrinda K, Weinstein R, Fleischer AS (2011) A direct comparison of three different
material enhancement methods on the transient thermal response of paraffin phase change
material exposed to high heat fluxes. Int J Therm Sci 50:1639–1647
6. Zhao CY, Lu W, Tian Y (2010) Heat transfer enhancement for thermal energy storage using
metal foams embedded within phase change materials (PCMS). Sol Energy 84:1402–1412
7. Xiao X, Zhang P, Li M (2014) Effective thermal conductivity of open cell metal foams
impregnated with pure paraffin for latent heat storage. Int J Therm Sci 81:94–105
8. Krishnan S, Murthy JY, Garimella SV (2007) Analysis of solid-liquid phase change under
pulsed heating. J Heat Transf 129:395–400
9. Yang Z, Garimella SV (2010) Melting of phase change materials with volume change in
metal foams. J Heat Transf 132:062301
72 4  PCM Design Issues

10. Fukai J, Kanou M, Kodama Y, Miyatake O (2000) Thermal conductivity enhancement of


energy storage media using carbon fibers. Energy Convers Manag 41:1543–1556
11. Warzoha R, Sanusi O, McManus B, Fleischer AS (2013) Development of methods to fully
saturate carbon foam with paraffin wax phase change material for energy storage. J Sol
Energy Eng 135:021007
12. Song J, Guo Q, Zhong Y, Gao X, Feng Z, Fan Z, Shi J, Liu L (2012) Thermophysical proper-
ties f high density graphite foams and their paraffin composites. New Carbon Mater 27:27–34
13. Zhong Y, Guo Q, Li S, Shi J, Liu L (2010) Heat transfer enhancement of paraffin wax using
graphite foam for thermal energy storage. Sol Energy Mater Sol Cells 94:1011–1014
14. Balandin AA, Ghosh S, Bao W, Calizo I, Teweldebrhan D, Miao F, Lau CN (2008) Superior
thermal conductivity of single layer graphene. Nanoletters 8:902–907
15. Zhong Y, Zhou M, Huang F, Lin T, Wan D (2013) Effect of graphene aerogel on thermal
behavior of phase change materials for thermal management. Sol Energy Mater Sol Cells
113:195–200
16. Marconnet AM, Panzer MA, Goodsen KE (2013) Thermal conduction phenomena in carbon
nanotubes and related nanostructured materials. Rev Mod Phys 85:1295–1326
17. Elgafy A, Lafdi K (2005) Effect of carbon nanofiber additives on thermal behavior of phase
change materials. Carbon 43:3067–3074
18. Wang J, Xie H, Xin Z (2009) Thermal properties of paraffin based composites containing
multi-walled carbon nanotubes. Thermochim Acta 488:39–42
19. Chintakrinda K, Warzoha R, Weinstein RD, Fleischer AS (2012) Quantification of the impact
of embedded graphite nanofibers on the transient thermal response of paraffin phase change
material exposed to high heat fluxes. J Heat Transf 134:071901
20. Sanusi O, Warzoha R, Fleischer AS (2011) Energy storage and solidification enhancements
of phase change materials embedded with nanofibers. Int J Heat Mass Transf 54:4436–4442
21. Fan L, Fang X, Wang X, Zeng Y, Xiao Y, Yu Z, Xu X, Hu Y, Cen K (2013) Effects of vari-
ous carbon nanofillers on the thermal conductivity and energy storage properties of paraffin
based nanocomposite phase change materials. Appl Energy 110:163–172
22. Kim S, Drazal LT (2009) High latent heat storage and high thermal conductivity phase
change materials using exfoliated graphite nanoplatelets. Sol Energy Mater Sol Cells
93:136–142
23. Warzoha R, Fleischer AS (2014) Effect of graphene folding on thermal conduction in nano-
composite. ACS Appl Mater Interfaces 6:12868–12876
24. Gui X, Tang D, Linag S, Lin B, Yuan X (2012) Influence of void ratio on thermal perfor-
mance of heat pipe receiver. Int J Heat Fluid Flow 33:109–117
25. Ferrer G, Sole A, Barreneche C, Martorell I, Cabeze LF (2015) Corrosion of metal containers
for use in PCM energy storage. Renew Energy 76:465–469
26. Zheng Y, Zhao W, Sabol J, Tuzla K, Neti S, Oztekin A, Chen JC (2013) Encapsulated phase
change materials for energy storage- characterization by calorimeter. Sol Energy 87:117–126
27. Zukowski M (2007) Experimental study of short term thermal energy storage unit based
on enclosed phase change material in polyethylene film bag. Energy Convers Manag
48:166–173
28. Esakkimuthu S, Hassabou AH, Palaniappan C, Spinnler M, Blumenberg J, Velraj R (2013)
Experimental investigation on phase change material based thermal storage system for solar
air heating applications. Sol Energy 88:144–153
29. Tyagi VV, Kaushik SC, Tyagi SK, Akiyama T (2011) Development of phase change materi-
als based microencapsulated technology for buildings: a review. Renew Sustain Energy Rev
15:1373–1391
30. Qiu X, Li W, Sing G, Chu X, Tang G (2012) Fabrication and characterization of microencap-
sulated n-octadecane with different cross-linked methymethacrylate-based polymer shells.
Sol Energy Mater Sol Cells 98:283–293
31. Li W, Song G, Li S, Yao Y, Tang G (2014) Preparation and characterization of novel microP-
CMs with hybrid shells via the polymerization of two alkoxy silanes. Energy 70:298–306
References 73

32. Sari A, Alkan C, Karaipekli A, Uzun O (2009) Microencapsulated n-octadecane as phase


change material for thermal energy storage. Sol Energy 83:1757–1763
33. Latibari S, Mehrali M, Mehrali M, Mahlia T, Metselaar H (2013) Synthesis, characteriza-
tion and thermal properties of nanoencapsulated phase change materials via sol-gel method.
Energy 61:664–672
34. Zhang GH, Bon SAF, Zhao CY (2012) Synthesis characterization and thermal properties
of novel nanoencapsulated phase change materials for thermal energy storage. Sol Energy
86:1149–1154
35. Kwon HJ, Cheong IW, Kim JH (2010) Preparation of n-octadecane nanocapsules by using
interfacial redox initiation in miniemulsion polymerization. Macromol Res 18:923–926
36. Ehid R, Fleischer AS (2012) Development and characterization of paraffin-based shape stabi-
lized energy storage materials. Energy Convers Manag 53:84–91
37. Ehid R, Weinstein RD, Fleischer AS (2012) The shape stabilization of paraffin phase change
material to reduce graphite nanofibers settling during the phase change process. Energy
Convers Manag 57:60–67
38. Kenisarin MM, Kenisarina KM (2012) Form-stable phase change materials for thermal
energy storage. Renew Sustain Energy Rev 16:1999–2040
39. Mehrali M, Latibari ST, Mehrali M, Metselaar HSC, Silakhori M (2013) Shape stabilized
phase change materials with high thermal conductivity based on paraffin graphene oxide
composite. Energy Convers Manag 67:275–282
Chapter 5
Fundamental Thermal Analysis

5.1 Introduction to the Analysis of Solid–Liquid


Phase Change

In order to fully understand the physics of phase change materials, it is impor-


tant to consider the mathematics that describe melting and solidification. We will
start with energy storage potential because that is much simpler to quantify than
the dynamic melting and solidification process. The energy storage potential of a
material is quantified using the latent heat of the material. As discussed in Chap. 3,
the latent heat of fusion, Lf, is the measure of how much energy can be stored in a
specified mass of material as it transitions from solid to liquid and is given in units
of kJ/kg. The energy storage potential of a material as it melts is found by multi-
plying the mass of material by the latent heat of fusion:
Estored = mLf (5.1)
However, whether the full energy storage potential of a material will be utilized,
and the rate at which the energy is stored, is dependent on the rate of heat transfer
into and out of the PCM mass. The rate of heat transfer is dependent on the tem-
perature differential between the melt point and the heat source, the boundary con-
ditions of the material and the initial temperature condition of the PCM.
Heating a material through the melting process generally occurs over three distinct
steps. Initially the material is raised uniformly in temperature from its initial tem-
perature to the melt point through sensible heating. Once the melt point is reached
and as the heating process continues, the material no longer increases in temperature
with the addition of heat, but now changes phase through the latent heating cycle.
During this time the material transitions from solid to liquid and the melt front, which
marks the transition point from solid to liquid, progresses through the material. As the

© The Author(s) 2015 75


A.S. Fleischer, Thermal Energy Storage Using Phase Change Materials,
SpringerBriefs in Thermal Engineering and Applied Science,
DOI 10.1007/978-3-319-20922-7_5
76 5  Fundamental Thermal Analysis

melt front crosses the material and the material becomes fully melted phase, further
addition of heat raises the temperature of the now liquid material through sensible
heating.
This process can be quite complex in reality. The motion of the melt front
through the PCM depends on the interaction of the boundary conditions at its
edges. The melt front can speed up or slow down depending on the rate of heat
added and lost through the boundaries. This creates a situation where the location
of the melt front at any point in time is not known in advance but must be solved
for. As the boundaries interact, the temperature distribution within the PCM takes
on a three-dimensionality, creating a situation where the melt front does not sim-
ply move uniformly from one side to the other across the material, but can instead
move diagonally. The initiation of natural convection with the melt phase leads to
further complexity in the melt front motion and the mathematics of the situation.
The use of PCMs that are blended materials, or which feature a melt range
rather than a sharp melt temperature leads to the development of a broad “mushy
zone” in which the melt front motion is indistinct. In these types of materials, the
material melts slowly over a wider temperature range. Due to the broader melt
zone the material slowly softens throughout and regions of liquid and solid phases
can be found interspersed within the mass of PCM.
For these reasons, it is challenging to develop exact analytical solutions to the
melting and solidification processes which are widely applicable. Instead these
problems, which are known as moving boundary problems, are only solved analyti-
cally for situations in which the physics can be simplified with various assumptions.

5.2 Stefan Problem

Partial differential boundary value problems with moving boundaries resulting


from melting and solidification are often referred to as Stefan problems in recogni-
tion of Josef Stefan’s work on this class of problem in the 1890s [1, 2]. Although
these problems were first worked on by Lamé and Clapeyron in 1831 [3], it
is Stefan’s pioneering work on the subject of ice formation that ended up being
immortalized in both the name of the class of problem, and in the nondimensional
Stefan number. The Stefan number, Ste, is the ratio of specific heat to latent heat
of fusion as given in Eq. (5.2), where ΔT is the difference in temperature between
the heat source (or sink) and the melting point of the material.
cp T
Ste = (5.2)
Lf
Materials with high Stefan numbers are dominated by the sensible heating process
and are generally unsuitable for use as PCMs. Most PCMs will have high latent
heats which allow them to store greater amounts of energy, and as such will have
lower Stefan numbers.
The classical Stefan problem solves for the temperature distribution in a homoge-
neous medium which is undergoing a change of phase. The boundary condition between
5.2  Stefan Problem 77

Fig. 5.1  Classical Stefan problem formulation with solidification from left to right

the solid and liquid phase moves with time and as such these problems are an example
of a free boundary problem. The coupling of the melt front motion to the solution of
the PDE makes the analytical solution difficult to achieve, and as such exact solutions
to these problems exist only for specific conditions. The classical assumptions made
in order to solve a Stefan problem are that the situation is one-dimensional and semi-
infinite. The material undergoing the phase change is assumed to have constant prop-
erties with a constant initial temperature and a constant temperature boundary on one
side that drives the phase change process. The melt front is assumed to exist as a sharp
interface between the phases. A standard Stefan problem is shown in Fig. 5.1.
Figure 5.1 illustrates a solidification problem. The situation is 1-D with cooling
at the left hand boundary where the boundary temperature Tb is lower than the melt
point of the material, Tm, which leads to solidification at the left edge which then
drives a solidification front across the material from left to right. The solidification
front location changes with time and its location at any point is denoted as S(t).
The Stefan problem formulation ignores convection in the melt phase, and the
governing equations for both the solid and liquid phases are developed from the
heat diffusion equation as given in Eq. 5.3,
     
  ∂T ∂ ∂T ∂ ∂T ∂ ∂T
ρcp = k + k + k + q̇
∂t ∂x ∂x ∂y ∂y ∂z ∂z (5.3)
Reducing Eq. 5.3 to a one-dimensional situation in x as illustrated in Fig. 5.1, and
removing the internal heat generation terms yields,
 
  ∂T ∂ ∂T
ρcp = k
∂t ∂x ∂x (5.4)
Equation 5.4 is then applied separately to the solid phase and to the liquid phase
which are then denoted using the subscripts s and l as seen in Eqs. 5.5 and 5.6.
The conductivity is assumed to be constant and pulled out of the differential.
 ∂Ts ∂ 2 Ts
(5.5)

ρcp s ∂t
= ks 2
∂x
  ∂Tl ∂ 2 Tl
ρcp l ∂t
= kl (5.6)
∂x 2
78 5  Fundamental Thermal Analysis

Equation  5.5 is applied in the region where 0 ≤  x  ≤  S(t) and Eq. 5.6 is applied
in the region where S(t)  ≤  x  ≤  ∞. The two phases are coupled at the interface
S(t) through the Stefan condition which is given by Eqs. 5.7 and 5.8. Equation 5.7
is the condition which couples the conduction from the solid phase to conduc-
tion in the liquid phase and to the motion of the solidification front at the inter-
face S(t). Equation 5.8 represents the continuity in temperature across the interface
where the temperature of the solid phase and the temperature of the liquid phase
are equal to the solidification temperature. Both equations are applied at location
x = S(t). The velocity of the solidification front, u, is given by Eq. 5.9.
∂Ts ∂Tl ∂S
ks − kl = Lf ρ s (5.7)
∂x ∂x ∂t

T (x, t)l = Ts (x, t) = Tm (5.8)

∂S
u= (5.9)
∂t
Finally, the boundary and initial conditions need to be identified for the problem.
The initial condition (Eq. 5.10) sets up a one region problem where the entire
region is originally constant temperature at the solidification/melt temperature.
The left boundary condition (Eq. 5.11) is a constant temperature condition with a
temperature below the solidification temperature and the right boundary condition
(Eq.  5.12) is that of a semi-infinite condition. The problem thus develops with a
solidification front starting at x = 0 at t = 0 and moving in the positive x direction.
The temperature within the liquid at any point in time remains at Tm as the initial
temperature is Tm and the boundary condition for the liquid is semi-infinite at the
right hand side. This formulation is known as a one region problem due to this
constant temperature in the liquid phase.
Tl (x, 0) = Tm (5.10)

Ts (0, t) = Tb < Tm (5.11)

Tl (x, t)x→∞ = Tm (5.12)


This formulation of the Stefan problem can classically solved using similarity
methods with a similarity variable as defined in Eq. 5.13. The transformation of
the governing equations is given in Eq. 5.14 where the temperature is defined as
in Eq. 5.15. Solving this transformed equation yields a transcendental equation as
given in Eq. 5.16 where λ is as defined in Eq. 5.17.
x
η= √ (5.13)
2 αt

d2θ dθ
2
+ 2η =0 (5.14)
dη dη
5.2  Stefan Problem 79

θ (η) = T (η) − Tb (5.15)

2 1
e erf() = √ Ste (5.16)
π

S(t)
= √ (5.17)
2 αt
The transcendental equation can be solved for λ for a given Ste as defined in
Eq.  5.18 to find the location of the interface S(t) at any given point in time
(Eq. 5.19) and the temperature distribution within the solid (Eq. 5.20). Ste should
be constant for a given material and boundary condition.
cp (Tm − Tb )
Ste = (5.18)
Lf

S(t) = 2 αt (5.19)

θ (η) erf(η)
= (5.20)
Tm − T b erf()
Finally, the formulation in Eq. 5.20 can be rearranged and the definitions in
Eqs.  5.13, 5.15 and 5.17 used to substitute for η and λ to yield a slightly more
user friendly form for the temperature distribution in the solid as given in Eq. 5.21.
Then, using Fourier’s law (Eq. 5.22), the heat flux can found as per Eq. 5.23.
 
erf 2√xαt
Ts (x, t) = Tb + (Tm − Tb )   (5.21)
erf 2S(t)

αt

 
′′ dT
q =k (5.22)
dx x=0

k(Tm − Tb ) −0.5
q′′ = √ t (5.23)
π α erf()
By examining Eq. 5.23 it can be seen that the heat flux at the surface decreases as
time increases, which is due to the increasing thermal resistance at the surface due
to the thickening solid layer. This has been discussed in previous chapters, particu-
larly with respect to the need to enhance the thermal conductivity of the PCMs to
improve performance during the conduction dominated solidification. The math-
ematical basis for this is clearly seen here.
Of course, this formulation is simplified as noted earlier. The behavior
described through this formulation is more accurate for the conduction domi-
nated solidification process than it is for the melt process during which natural
80 5  Fundamental Thermal Analysis

Fig. 5.2  Melting problem formulation with natural convection in the melt

convection is induced. When natural convection is induced, the motion of the


melt front becomes two-dimensional as seen in Fig. 5.2, or even three-dimen-
sional depending on the interaction of the boundary conditions. This solution is
also limited to the constant temperature, semi-infinite boundary conditions. There
are some additional formulations of the exact solution available in the literature
for both melting and solidification cases with various boundary conditions [4–6],
some of which include the impact of nanoparticles on the solution [7].

5.3 Advanced PCM Analysis

For more complex PCM melting and solidification conditions, particularly ones in
which natural convection occurs in the melt phase, it is more common to model
the physics of the situation using computational models than to attempt to solve
for an exact solution. Various computational techniques can be used to success-
fully capture the two and even three dimensional nature of the melt process in a
pure PCM.
A common way to approach the modeling of the phase change process is to
use what is known as an enthalpy-porosity technique. In this method, instead of
a sharp interface at the melt front, the melt front region is modeled as a porous
medium whose porosity is a function of the liquid fraction. This creates a small
“mushy zone” at the interface. The momentum conservation equations then are
modified to include a flow resistance term which is a function of the porosity. The
source term varies smoothly across the mushy zone such that it ensures normal
viscous flow in the fully liquid region, allowing for convection in the melt, but
goes to zero flow in the fully solid region [8].
This method was used by Pal and Joshi [9] who accurately captured the melt
process in a vertically heated container containing pure triacontane, in an arrange-
ment similar to that shown in Fig. 5.2. The melt front was accurately predicted
to take on the shape of a diagonal from bottom left to top right due to the nat-
ural convection currents within the system. The code was validated against both
experimental data and imaging of the melt process which plainly showed the melt
5.3  Advanced PCM Analysis 81

front progressing as predicted. Models can be developed with varying degrees of


complexity and models which includes both natural convection in the melt phase
and the effect of volumetric expansion/contractions through the melt/solidification
transitions was developed by Shatikian et al. [10, 11].
Models such as these allow one to accurately track the progression of the melt
front and to identify the melt fraction at any given point in time. This allows
the determination of how much of the energy storage capability is being uti-
lized at any given point in time for an application. This data can be nondimen-
sionalized for wider applicability using the Stefan number (Eq. 5.18), the Fourier
number (Eq. 5.24) and the Rayleigh number in the melt (Eq. 5.25). Many com-
mercially available computational codes can accurately capture the melt and solid-
ification process if the PCM properties and boundary conditions can be accurately
represented.
αt
Fo = (5.24)
L2

gβ(Tb − Tm )L 3
Ra = (5.25)
να
The representation of the PCM properties in the models can be problematic how-
ever. The variation in the properties of the material as it transitions through the
melt process can affect the progression of the melt front and the influence of these
properties must be accurately captured in the model. The addition of materials
to improve the thermal diffusivity as discussed in Chap. 4 leads to complexity in
modeling these properties. The impact of the inclusions, whether foams, fibers or
nanoparticles, on the properties can be difficult to capture accurately.
There are many different analytical models available to try to capture the effect
of foams on the PCM thermal conductivity. Many of these models assume ther-
mal equilibrium between the foam and the PCM and develop “effective” thermal
properties by treating the matrix and PCM as a homogeneous mixture with aver-
aged properties. These are known as one-temperature models and are based on the
porosity of the foam. For instance, in a simplified version, the volumetric energy
storage can be represented using Eq. 5.26 in either the solid or liquid phase and by
Eq. 5.27 during the phase change process while the effective thermal conductivity
can be found using Eq. 5.28 where ε is the foam porosity.
(ρcP )eff = ε(ρcP )PCM + (1 − ε)(ρcP )foam (5.26)
 
ρLf PCM
(ρcP )eff = ε(ρcP )PCM + (1 − ε)(ρcP )foam + (5.27)
∆T

keff = εkPCM + (1 − ε)kfoam (5.28)


A number of more complex expressions with greater accuracy have been devel-
oped for the effective thermal conductivity of foams and a number are listed below
(Eqs.  5.29–5.34) [12–14] including the well-known Maxwell-Garnett effective
82 5  Fundamental Thermal Analysis

medium theory in which the matrix is considered to be the foam and the inclu-
sions are considered to the PCM (Eq. 5.29). The formulation in Eq. 5.32 requires
Eq. 5.33 where r is defined as the ratio of the ½ the foam ligament thickness to ½
the thickness of the joints between ligaments.
kPCM (1 + 2ε) + 2kfoam (1 − ε)
keff = kfoam (5.29)
kPCM (1 − ε) + kfoam (2 + ε)

(1 − ε)
keff = kfoam [13] (5.30)
3
 1/3
keff
keff = kfoam − (kfoam − kPCM )ε [14] (5.31)
k PCM
  � � � �
b b
r (1 − r)
√ Lf Lf
�� �
keff = 2 3  � �� �+ 2b(kfoam −kPCM )
b kfoam −kPCM kPCM +
kPCM + 1 + Lf 3 3Lf
−1

3 b
2 − Lf 
+ � � 
 [15] (5.32)
4rb
kPCM + √ (kfoam − kPCM )
3 3Lf

   
−r + r 2 + √2 (1 − ε) 2 − r 1 + √4
b 3 3
=    [15] (5.33)
Lf 2
2 − r 1 + √4
3 3

    
kPCM + π 1−ε 1−ε kPCM + 1−ε
3π − 3π (kfoam − kPCM ) 3 (kfoam − kPCM )
keff =     [16]
kPCM + 3 4 1−ε 1−ε
3π (1 − ε) + π 3π − (1 − ε) (kfoam − kPCM )
(5.34)

As noted earlier, Eqs. 5.26–5.34 all assume a one temperature model where the
PCM and the foam are in thermal equilibrium. However, if the thermal diffusivity
of the foam is much higher than that of the PCM then the system is more accu-
rately represented by a two-temperature model [15]. In a two-temperature model
the foam and the PCM are assumed to be at different temperatures and heat trans-
fer between the foam and the PCM must be accounted for. For these models, the
conductivity is not developed as an effective conductivity for the combined mate-
rial, but is instead modeled separately for the foam and the PCM. It is common in
this case to use the effective conductivity models with either kPM or kfoam set to
zero to develop a model for the thermal conductivity as a function of porosity. An
example of this is shown in Eqs. 5.35 and 5.36 where Eq. 5.34 is used as the start-
ing basis [15].
5.3  Advanced PCM Analysis 83

   
1−ε 1−ε 1−ε
π 3π − 3π 3 kfoam
keff, foamonly =   [16]
(5.35)

4 1−ε 1−ε
3 3π (1 − ε) + π 3π − (1 − ε)

    
1−ε 1−ε
kPCM − π 3π − 3π (kPCM ) kPCM − 1−ε 3 kPCM
keff, PCM =     (5.36)
[16]
kPCM + 43 1−ε
3π (1 − ε) + π 1−ε
3π − (1 − ε) kPCM

When dealing with nanoparticle enhanced PCMs rather than foams, many of
the same basic models can be used with slight modifications. For instance, the
Maxwell-Garnett effective medium theory can be used, but in this case the PCM
will play the role of the matrix, the nanoparticles are the inclusions and the vol-
ume percent loading level of nanoparticles in the composite is represented by ϕ.
kNP (1 + 2ϕ) + 2kPCM (1 − ϕ)
keff = kPCM (5.37)
kNP (1 − ϕ) + kPCM (2 + ϕ)
There are several other models available for nanoparticle laden materials, most based
on variations of the effective medium theory. The models for nanoparticles generally
assume that the nanoparticles are completely separated from one another and homog-
enously distributed throughout the PCM, or that all the nanoparticles are connected
and are fully percolated throughout the PCM [16]. The upper bound of the thermal
conductivity is represented by the fully percolating model as given in Eq. 5.30 and
the lower bound of the thermal conductivity by the fully dilute model in Eq. 5.31.
Equation 5.30 is known as the parallel model and Eq. 5.31 as the series model [17].
keff = (1 − ϕ)kPCM + ϕkNP (5.38)

1
keff = (1−ϕ) ϕ (5.39)
kPCM + kNP

More complex representations of the effective thermal conductivity for PCM with
embedded nanoparticles have been developed which include the effects of nano-
particle shape (spherical or ellipsoidal), nanoparticle orientation, nanoparticle
aspect ratio and nanoparticle—matrix and/or nanoparticle-nanoparticle thermal
boundary resistance. Equation 5.32 was developed for noncontacting spherical
particles and includes the effects of Kapitza (boundary) resistance between the
nanoparticles and the PCM [18]. The parameter β is the ratio of the Kapitza radius
of the nanoparticle to the geometric radius of the nanoparticle. For a perfect inter-
face with no resistance between phases, β will be zero. Equations for noncontact-
ing ellipsoidal particles are also available [18].
kNP (1 + 2β) + 2kPCM + 2ϕ(kNP (1 − β) − kPCM )
keff = kPCM [19] (5.40)
kNP (1 + 2β) + 2kPCM − ϕ(kNP (1 − β) − kPCM )
84 5  Fundamental Thermal Analysis

The thermal conductivity of percolating matrices of nanoparticles is given by


Eq.  5.33 [19]. In Eq. 5.33, the parameter ϕcritical represents the volume fraction
at which the system reaches percolation, which may be a function of nanopar-
ticle aspect ratio. The factor k0 represents the conductivity of nanoparticles and
includes the effect of interparticle contact resistance and the topology of the per-
colation cluster. The exponent t depends on the dimensionality of the system (2D
or 3D) and the geometry of the nanoparticle. The exponent t can range as high as
2.0 for spheres in a 3D system to as low as 1.1 for ellipsoids in a 2D system. The
most common range for t is 1.1–1.6 [19]. An alternate derivation for the effective
thermal conductivity of percolating matrices is developed in [20].
keff = k0 (ϕ − ϕcritical )t [20] (5.41)
Despite the complexity of these formulations, the accuracy of the results are quite
high. The development of models such as these to predict the transient response
and energy storage performance of PCMs in application is a great asset to the ther-
mal engineer wishing to use PCMs in their designs.

References

1. Stefan J (1889) Über die Theorie des Eisbildung, insbesondere _Über die Eisbildung in
Polarmeere. S B Wien Akad Mat Natur 98:965–983
2. Stefan J (1891) Uber die theorie der eisbildung, inbesondere uber die eisbildung im polar-
meer. Annalen der Physik und Chemie 42:269–286
3. Lame G, Clapeyron BPE (1831) Memoire sur la solidification par refroidissment d’un globe
solid. Ann Chem Phys 250–60
4. Hu H, Argyropoulos SA (1996) Mathematical modelling of solidification and melting: a
review. Model Simul Mater Sci Eng 4:371–396
5. Krishnan S, Garimella S (2004) Thermal management of transient power spikes in electron-
ics—phase change energy storage or copper heat sinks? J Electr Pack 126:308–316
6. Lu TJ (2000) Thermal management of high power electronics with phase change cooling. Int
J Heat Mass Transf 43:2245–2256
7. El Hasadi YMF, Khodadadi JM (2013) One dimensional Stefan problem formulation for
solidification of nanostructure enhanced phase change material. Int J Heat Mass Transf
67:202–213
8. Nayak KC, Saha SK, Srinivasan K, Dutta P (2006) A numerical model for heat sinks
with phase change materials and thermal conductivity enhancers. Int J Heat Mass Transf
49:1833–1844
9. Pal D, Joshi YK (2001) Melting in a side heated tall enclosure by a uniformly dissipating
heat source. Int J Heat Mass Transf 44:375–387
10. Shatikian V, Ziskind G, Letan R (2005) Numerical investigation of a PCM based heat sink
with internal fins. Int J Heat Mass Transf 48:3689–3706
11. Shatikian V, Ziskind G, Letan R (2008) Numerical investigation of a PCM-based heat sink
with internal fins: constant heat flux. Int J Heat Mass Transf 51:1488–1493
12. Lemlich R (1978) A theory for the limiting conductivity of polyhedral foam at low density. J
Colloid Interf Sci 6:107–110
13. Weaver JA, Viskanta R (1986) Freezing of liquid-saturated porous media. J Heat Transf
108:654–659
References 85

14. Calmidi VV, Mahajan RL (1999) The effective thermal conductivity of high porosity fibrous
metal foams. J Heat Transf 121:466–471
15. Mesalhy O, Lafdi K, Elgafy A, Bowman K (2005) Numerical study for enhancing the ther-
mal conductivity of phase change material storage using high thermal conductivity porous
matrix. Energy Convers Manag 46:847–867
16. Warzoha R, Weigand R, Fleischer AS (2015) Temperature-dependent thermal properties of
a paraffin phase change material embedded with herringbone style graphite nanofibers. Appl
Energy 137:716–725
17. Han Z, Fina A (2011) Thermal conductivity of carbon nanotubes and their polymer nano-
composites: a review. Prog Polym Sci 36:914–944
18. Nan CW, Birringer R, Clarke DR, Gleiter H (1997) Effective thermal conductivity of particu-
late composites with interfacial thermal resistance. J Appl Phys 81:6692–6699
19. Foygel M, Morris R, Anez D, French S, Sobolev V (2005) Theoretical and computational
studies of carbon nanotube composites and suspensions: electrical and thermal conductivity.
Phys Rev B 71:104201
20. Wemhoff AP (2013) Thermal conductivity predictions of composites containing percolated
networks of uniform cylindrical inclusions. Int J Heat Mass Transf 62:255–262
Chapter 6
Future Directions

6.1 Development Needs

While phase change materials are being explored for all kinds of applications from
energy storage for large scale solar energy systems to the thermal management of
portable electronics, and from energy-efficient building materials to thermal pro-
tection for space based systems, it is clear that further development is needed to
ensure their successful entry into the marketplace. Most of this development is
focused on the design of new PCMs with enhanced thermal properties, because it
is apparent that no existing material features all the desirable characteristics of an
ideal PCM.
Based on our study of PCMs to date, an ideal PCM would feature the following
characteristics:
• High latent heat
• High specific heat
• High thermal conductivity
• Small density change through the phase transition
• Low or no supercooling tendency
• Chemical stability
• Non flammable
• Compatible with most materials
• Low cost
While many of the materials available now possess some of these features, none
of them possess all of them. Thus, materials development work is ongoing to cre-
ate new, improved options for energy storage materials. Development of high tem-
perature PCMs with transition points above 500 °C are particularly important as
enabling technologies for higher efficiency CSP plants.

© The Author(s) 2015 87


A.S. Fleischer, Thermal Energy Storage Using Phase Change Materials,
SpringerBriefs in Thermal Engineering and Applied Science,
DOI 10.1007/978-3-319-20922-7_6
88 6  Future Directions

6.2 Conductivity Enhancement

The increase of thermal conductivity continues to be of interest for most PCM


applications. Recent work in this area is focusing on the impact of single or few
layer graphene on the effective thermal conductivity. Interestingly, the use of
nanoscale graphene platelets in PCM composites is yielding effective thermal con-
ductivities far above those seen with MWCNT, SWCNT, or GNF inclusions. For
example, Warzoha and Fleischer [1] studied the increase in thermal conductivity
resulting from the addition of either MWCNT, Alumina, TiO2 or XNGP at loading
levels of 20 vol% in a base paraffin. At these loading levels, the increases in meas-
ured thermal conductivity were high for all composites. The thermal conductivity
of the PCM/TiO2 composite increased 76 %, the PCM/Alumina increased 216 %,
and the PCM/MCNT increased 832 %. However, all of these pale in comparison
to the effect of adding 20 vol% exfoliated graphene platelets, which increased the
thermal conductivity 2800 %, from about 0.2 W/m K to about 5.6 W/m K. It is
clear that the xGNP affects the structure of the paraffin in a way that is unique and
different from the other types of carbon and metal based nanoparticles.
The nature of this enhancement was studied by Babaei et al. [2] who used
molecular dynamics simulations to study the behavior of paraffin molecules in the
presence of graphene nanoparticles. Their simulations used the non-equilibrium
molecular dynamics (NEMD) method to study n-octadecane and n-octadecane
with embedded MWCNT and graphene. A heat flux was imposed through the sim-
ulation space by adding heat to the molecules in a heat source on one side of the
space and extracting the same amount of heat from molecules in a heat sink on
the opposite side. The steady-state results were used with Fourier’s law to extract
the thermal conductivity. The resulting model showed that the thermal conductiv-
ity exhibited a strong dependence on the relative alignment of the PCM and the
nanoinclusions. The graphene/PCM composite featured an “enhanced crystalliza-
tion” in the solid phase due to directed ordering of PCM molecules in near prox-
imity to the graphene molecules. This ordered crystallization effect did not occur
with the CNT inclusions, and the enhanced thermal conductivity of the graphene
composites is thought to depend in part at least on this crystallinity because it is
known that higher degrees of crystallinity lead to higher thermal conductivities.
Thus it appears that graphene based nanoparticles interact with the PCM mole-
cules in a way that affects their basic alignment, and more work is needed to deter-
mine the full impact of this effect.

6.3 Specific Heat Enhancement

Improvements in specific heat are desirable in next generation PCMs, in order to


increase the amount of energy which can be stored during the sensible heating
periods of operation. Several studies have showed that the enhanced crystallinity
6.3  Specific Heat Enhancement 89

of PCMs in certain composites can also increase the specific heat capacity of the
material [3, 4]. Rather than study graphene in paraffin, these studies considered the
use of silicon dioxide nanoparticles (2–26 mm diameter) in molten salt eutectics.
In the first study [3], 26 nm SiO2 particles were added at 1 wt% into a eutectic
of alkali chloride salts (BaCl2, NaCl, CaCl2, LiCl) with a melt point of 378 °C.
Repeated measurements using the DSC method showed an increase in specific heat
capacity of the material of over 14 % with the SiO2 addition. The reasons for this
increase were theorized to be a combination of the high specific heat of nanosilica
compared to bulk silica values, the interfacial thermal resistance introduced at the
nanoparticle-PCM interface, and “layering” of liquid PCM molecules at the silicon
particle surface.
This effect was studied in greater detail using SiO2 particles 2–20 nm in diameter
embedded in a molten salt eutectic of Li2CO3–K2CO3 [4]. DSC measurements showed
an increase in specific heat of 38–54 % for the solid phase of these composite materi-
als and 118–124 % for the liquid phase of the composites over the pure eutectic. SEM
imaging of the composites revealed that substructures which resembled a woven cloth
pattern had formed in the molten salt at the edges of the nanoparticles. This “woven”
structure creates a denser layer at the particle edge which appeared to mimic the under-
lying crystal lattice [4]. It is thought that the formation of this layer is driving the
increase in specific heat.

6.4 Latent Heat Enhancement

As energy storage through the phase change process is directly proportional to the
latent heat of the material, it is always of interest to develop materials with higher
latent heat capacity. This allows either the storage of more energy within the same
material mass, or the use of reduced mass levels for a constant energy storage
need.
As with thermal conductivity and specific heat, it appears that if the crystal-
linity of the PCM can be increased, higher latent heats will result. Warzoha and
Fleischer [1] studied the increase in latent heat resulting from the addition of
either MWCNT, Alumina, TiO2 or XNGP to a base paraffin at loading levels of
20 vol%. In most cases, it is difficult to increase the latent heat of a material by
adding nanoinclusions. As the nanoparticle do not change phase, the latent heat
of the composite typically decreases by the loading level of the nanoparticles.
This value of latent heat of the composite, which is typically measured by DSC is
known as the “apparent phase change enthalpy” of the composite.
The impact of the nanoparticles on the latent heat of the PCM itself is a bit
harder to discern. In order to isolate the latent heat of the PCM, you have to cal-
culate the absolute phase change enthalpy which is the apparent phase change
enthalpy of the composite as measured by DSC divided by the PCM mass fraction.
This value can then be compared to the latent heat value for the PCM to ascer-
tain any impact of the nanoparticles. In this study [1], the absolute phase change
90 6  Future Directions

enthalpies of the 20 vol% MWCNT, Alumina and TiO2 composites where within
1 % of the latent heat of the base PCM, indicating no effect of the nanoparticles on
the latent heat as the measurement was within the accuracy of the DSC measure-
ment. However, the absolute phase change enthalpy of the 20 vol% graphene/PCM
composite was 13 % higher than the latent heat of the base PCM, once again sug-
gesting that graphene interacts with PCM in a unique manner, perhaps becoming
more crystalline across the faces of the graphene nanoparticles. As the latent heat
reflects the energy needed to change the existing crystal structure into a disordered
state, the greater the degree of crystallinity at the onset, the higher the latent heat
required.
Similar results were seen in by Li et al. [5] who studied “spongy” graphene in
docosane who found 2.5 % rise in apparent phase change enthalpy. Although they
did not provide results for absolute phase change enthalpy it would be expected to
be high since the overall latent heat of the composite increased.

6.5 Conclusions

While there is much work ongoing to develop PCMs with higher thermal conduc-
tivities, higher specific heats and higher latent heats, work is still needed on other
aspects of PCM performance. As new eutectic blends are developed, the physi-
cal and chemical stability of these blends needs to be guaranteed. The safety of
PCMs for certain applications is also an issue, as paraffins are flammable, and thus
have usage concerns when applied in wallboard and other building materials, and
in space based systems.
Finally, the cost of most PCMs is currently preventing wider application of
these materials. Pure research grade paraffins with well-defined thermal proper-
ties are too expensive for sustained usage in systems with short payback periods,
and the thermal properties of less expensive blends or eutectic mixtures are not yet
well defined.
As shown in this text, the potential applications for PCMs for energy storage
and for thermal management are vast, and as the development work addresses
these current concerns through the creation of next generation PCMs, these mate-
rials should find wide acceptance.

References

1. Warzoha R, Fleischer AS (2014) Improved heat recovery from paraffin-based phase change
materials due to the presence of percolating graphene networks. Int J Heat Mass Transf
79:324–333
2. Babaei H, Keblinski P, Khodadadi J (2013) Thermal conductivity enhancement of paraffins by
increasing the alignment of molecules through adding CNT/graphene. Int J Heat Mass Transf
58:209–216
References 91

3. Shin D, Banerjee D (2011) Enhancement of specific heat capacity of high temperature silica-
nanofluids synthesized in alkai chloride salt eutectics for solar thermal-energy storage applica-
tions. Int J Heat Mass Transf 54:1064–1070
4. Shin D, Banerjee D (2013) Enhanced specific heat capacity of nanomaterials synthesized by
dispersing silicia nanoparticles in eutectic mixtures. J Heat Transf 135:032801
5. Li JF, Lu W, Zeng YB, Luo ZP (2014) Simultaneous enhancement of latent heat and thermal
conductivity of docoasane based phase change materials in the presence of spongy graphene.
Sol Energy Mater Sol Cells 128:48–51
Index

A F
Advanced PCM analysis, 80 Form stable PCMs, 68
Advantages of PCM, 4
Apparent phase change enthalpy, 89
H
Heat exchanger designs, 24
B
Basic designs, 62
Brush type, 55 I
Building materials, 12 Inorganic PCM, 41, 43, 44
Introduction to PCM, 1
Issues of PCM design, 49
C
Cliff palace, 12
Concentrating solar power (CSP), 15, 16, L
18, 19, 21 Latent heat, 38, 40, 43–45
Conductivity enhancement, 88 Latent heat enhancement, 89
Latent heat of fusion, 1, 2
Layering, 89
D Liquid metals, 37
Design issues of PCM, 49
Development needs, 87
Domestic solar applications, 15 M
Melting, 75, 76, 80
Micelles, 67
E Microencapsulated PCM, 62, 64, 66
Effective thermal conductivity, 81, 83, 84 Macroscale carbon inclusions, 55
Energy storage, 7, 9, 12–14, 16–21, 24, 25, Metal alloy PCMs, 44
28, 30 Metallic inclusions, 52
Energy storage applications, 7 Metal PCMs, 44
Enhanced crystallization, 88 Micro-encapsulated PCMs, 66
Enhancement, 50 Micro-PCM, 67
European space agency (ESA), 31

© The Author(s) 2015 93


A.S. Fleischer, Thermal Energy Storage Using Phase Change Materials,
SpringerBriefs in Thermal Engineering and Applied Science,
DOI 10.1007/978-3-319-20922-7
94 Index

Moore’s law, 8 S
Mushy zone, 40, 76, 80 Shape-stabilized PCM, 68
Solar energy systems, 15
Solar power plants, 15
N Solidification, 75–81
Nanoscale carbon inclusions, 58 Solid–liquid phase change, 75
Nano-enhanced PCM, 83 Space systems, 29
NanoPCM, 66, 68 Specific heat, 38, 40, 42
Specific heat enhancement, 88
Stefan problem, 76–78
O
Organic PCM, 39–41
Outlast/silk, 28 T
Textile designs, 26
Thermal abatement, v
P Thermal analysis, 75
Packed bed designs, 21 Thermal capacitors, 7, 30
PCM containment, 62 Thermal conductivity, 50
PCM design, 50, 87 Thermal conductivity enhancement, 88
PCM development, 31, 50, 62, 84 Thermal energy storage, 3–5
PCM Implementation, 49 Thermal management, 3, 4
PCM properties, 37, 38, 41, 43–46 Thermal management of electronics, 8, 10, 27
PCM selection, 37 Types of PCMs, 37
PCM thermal conductivity, 49, 50, 60
Phase change, 75
Phase change basics, 1
Phase change materials (PCM), 1, 3–5,
7–15, 17–32
Pre-charged, 31

You might also like