Chapter 3

Phase Change Materials for Application in Energy-Efficient Buildings

B.P. Jelle1,2 and S.E. Kalnæs1,    1Norwegian University of Science and Technology (NTNU), Trondheim, Norway,    2SINTEF Building and Infrastructure, Trondheim, Norway

Abstract

Phase change materials (PCM) have received considerable attention over the last decade for use in latent heat thermal storage (LHTS) systems. PCMs give the ability to store passive solar and other heat gains as latent heat within a specific temperature range, leading to a reduction of energy usage, an increase in thermal comfort by smoothing out temperature fluctuations throughout the day, and a reduction and/or shift in peak loads. The interest around PCMs has been growing significantly over the last decade, both for new buildings and retrofitting of existing ones. Hence, several commercial products have arrived on the market with various areas of use in building applications. This study reviews commercial state-of-the-art products found on the market and shows some of the potential areas of use for PCMs in building applications. Examples of how PCMs can be integrated into buildings, and furthermore building materials and projects using PCMs that have already been realized, have also been reviewed. There seems to be a scarcity of data published on actual performance in real-life applications so far. However, many laboratory and full-scale experiments have shown positive results on energy savings. Furthermore, future research opportunities have been explored and challenges with the technology as of today have been discussed.

Keywords

Phase change material; PCM; energy; temperature; building; state-of-the-art; review

3.1 Introduction

In the world of today there is a growing energy demand, which presents challenges to the ever-increasing issue regarding climate change and energy supply. The world consumes large amounts of fossil fuels, driving climate change and emptying the supplies of fossil fuels more rapidly. One action that will benefit energy usage globally is to increase the energy efficiency of buildings. In the European Union, the building sector is a major energy consumer and accounts for around 40% of the total energy usage. Large parts of this energy usage are directly related to the heating and cooling of buildings (European Union, 2012). An alternative to meet the increasing energy demand and reduce the negative environmental impacts would be to reduce energy usage in combination with higher utilization of environmentally friendly renewable energy technologies. These may be well-known technologies such as hydropower, wind energy, and solar energy, or new emerging technologies such as ocean energy and enhanced geothermal systems, to mention a few. Building integration of some of these technologies may also be found to be beneficial, e.g., buildingintegrated photovoltaics (BIPV) (Jelle, 2016; Jelle and Breivik, 2012a,b; Jelle et al., 2012b).

Several new technologies are emerging to help realize the goal of reducing energy usage in buildings. Some of these technologies are related to thermal insulation materials applied in the building envelope (Baetens et al., 2010a,c, 2011; Brunner et al., 2014; Gao et al., 2013; Jelle, 2011a; Jelle et al., 2010, 2011b, 2014a, 2015a, 2015c; Kalnæs and Jelle, 2014; Koebel et al., 2012; Levy and Zayat, 2015; Sandberg et al., 2013; Tenpierik, 2009). Also part of the thermal building envelope, although not a thermal insulation material in itself, is the promising technology of phase change materials (PCM), which have received considerable attention since the first decade of the 21st century. PCMs utilize the principle of latent heat thermal storage (LHTS) to absorb energy in large quantities when there is a surplus, and release it when there is a deficit. Correct use of PCMs can reduce peak heating and cooling loads, i.e., reduce energy usage, and may also allow for smaller dimensions of technical equipment for heating and cooling. An added benefit is the ability to maintain a more comfortable indoor environment due to smaller temperature fluctuations. Since 2004, there have been written several reviews on the use of PCMs in buildings for thermal energy storage systems and indoor climate comfort purposes (Agyenim et al., 2010; Al-Saadi and Zhai, 2013; Baetens et al., 2010b; Cabeza et al., 2011; Cui et al., 2015; Farid et al., 2004; Giro-Paloma et al., 2016; Jurkowska and Szczygieł, 2016; Kalnæs and Jelle, 2015; Kenisarin and Mahkamov, 2007, 2016; Khadiran et al., 2015, Khudhair and Farid, 2004; Kibria et al., 2015; Konuklu et al., 2015; Kuznik et al., 2011a; Liu et al., 2015; Ma et al., 2016; Memon, 2014; Osterman et al., 2012; Pomianowski et al., 2013; Silva et al., 2016; Soares et al., 2013; Su et al., 2015; Tatsidjodoung et al., 2013; Thambidurai et al., 2015; Waqas and Din, 2013; Yang et al., 2015; Zalba et al., 2003; Zhang et al., 2016; Zhao and Zhang, 2011; Zhou et al., 2012; Zhu et al., 2009), thus clearly showing that the interest for PCMs is increasing worldwide.

PCMs may be applied for both new buildings and retrofitting of existing ones, and for building applications the possible areas where PCMs can be utilized are many. Some of the areas that have been studied to this day include ventilation systems, passive heating and cooling systems, floors, roofs, and wallboards. PCMs can also be incorporated directly into building materials such as concrete (Ling and Poon, 2013) and wallboards (Kuznik et al., 2008), enabling them to be applied in construction with minimal alterations to the original design.

The first objective of this study is to present a comprehensive state-of-the-art overview of different PCM manufacturers and products, including an evaluation of the effect and durability of these products. The second objective of this study is to bridge the gap from state-of-the-art PCMs to discuss and explore future research pathways and opportunities for PCMs. Furthermore, it is of interest to see how PCMs are tested with respect to lifetime performance in building applications, and then especially with respect to cycle testing for long-term stability, fire safety, and energy saving potential. These investigations may help to develop guidelines for a new testing scheme and point to future research possibilities. The work presented herein is based on the earlier study in a journal article by Kalnæs and Jelle (2015), wherein further details about specific PCM manufacturers and products may be found.

3.2 Phase Change Materials in General

3.2.1 General

Phase change materials (PCM) utilize the latent heat of phase change to control temperatures within a specific range. When the temperature rises above a certain point, the chemical bonds in the material will start to break up and the material will absorb the heat in an endothermic process where it changes state from solid to liquid. As the temperature drops, the material will give off energy and return to a solid state. The energy used to alter the phase of the material, given that the phase change temperature is around the desired comfort room temperature, will lead to a more stable and comfortable indoor climate, as well as cut peak cooling and heating loads (Baetens et al., 2010b). Hence, PCMs can provide an increase in heat storage capacity, especially in buildings with low thermal mass. The temperature range varies depending on the materials used as the PCM.

Schröder and Gawron (1981) summarized some of the desired properties that should be required from PCMs:

ent High heat of fusion per unit volume and unit weight, and high specific heat. This is desirable to gain more effect from latent heat storage with as small as possible volume of PCMs.

ent Phase change temperature suitably matched to the application. To gain the most out of PCMs the phase change temperature must be in accordance with the climate, location in the building or the type of system where the PCM is used.

ent Low vapor pressure at operational temperature. To avoid extra costs or danger of rupture because of pressure on the encapsulating material the vapor pressure should be as low as possible.

ent Chemical stability and low corrosion rate. Chemically stable materials will allow for PCMs to operate at the given temperature and with the given effect for a longer period and reduce the chances that the PCM reacts with materials that are in contact with it, i.e., increasing the lifetime of PCMs.

ent Not hazardous or poisonous. To be allowed in the building sector there can be no poisonous emissions during fire or if the encapsulation is ruptured during regular use. Production of the material should also not release dangerous emissions to the environment.

ent Highly inflammable. Strict building laws with regards to fire safety must also be fulfilled by PCMs.

ent Reproducible crystallization without degradation. Much the same as chemical stability, the reproduction of crystals over thousands of phase changes without degradation is vital to attain long lifetimes for PCMs.

ent Small degree of supercooling and a high rate of crystal growth. Supercooling will alter the temperature of the phase change. An attractive PCM should have an exact phase change temperature so that the phase change is predictable to allow a material to be selected correctly for optimal design.

ent Small volume change during solidification. A large volume change will mean that the encapsulation material has to allow for expansion of the PCM, thus decreasing the amount of PCM the encapsulation can hold.

ent High thermal conductivity. A high thermal conductivity will allow heat to disperse through or leave the material more rapidly, allowing the PCM to absorb or release heat at a higher rate.

ent Use materials that are abundant and cheap. To make the technology more attractive and possible to use at a large scale it is important that the materials to be used are abundant and cost-effective.

3.2.2 General Categorization of Phase Change Materials

There are several materials that can be used as PCMs. A common way to distinguish PCMs is by dividing them into organic, inorganic, and eutectic PCMs. These categories are further divided based on the various components of the PCMs (Fig. 3.1). Fig. 3.2 shows the difference in melting enthalpy and melting temperature for some of the most common materials used as PCMs.

image
Figure 3.1 General categorization of PCMs (Rathod and Banerjee, 2013).
image
Figure 3.2 Melting enthalpy versus melting temperature for various materials used in PCMs (Dieckmann, 2013).

3.2.2.1 Organic

Organic PCMs are divided into paraffins and non-paraffins. In general, organic PCMs do not suffer from phase segregation and crystallize with little or no supercooling (Rathod and Banerjee, 2013).

Paraffins are available in a large temperature range (Fig. 3.2), opening them up for use in various other areas besides building related applications. The latent heat is mass based, they show no signs of phase separation after repeated cycling through solid–liquid transitions, and have a low vapor pressure (Alkan, 2006). However, paraffins used as PCMs have some drawbacks. They have low thermal conductivity (around 0.2 W/(mK)), they are not compatible with plastic containers, and they are moderately flammable (Sharma et al., 2009).

Non-paraffins used as PCMs include fatty acids and their fatty acid esters and alcohols, glycols, etc. Fatty acids have received the most attention for use as PCMs in buildings. An extensive review on fatty acids used for PCM purposes has been written by Yuan et al. (2014). In this review, fatty acids and their esters and alcohols were reviewed for their potential as PCMs. The most interesting fatty acids for PCM purposes include lauric acid, myristic acid, palmitic acid, and stearic acid. As with paraffins, these also suffer from low thermal conductivity, ranging from 0.15 to 0.17 W/(mK).

Overall, organic PCMs have many qualities that make them suited for building applications. However, the fact that many organic PCMs are considered flammable is a crucial drawback that impacts the safety aspect of organic PCMs considerably when aimed at building applications.

3.2.2.2 Inorganic

Inorganic PCMs of interest consist of hydrated salts and metallics. For building applications, however, metallics are not within the desired temperature range, and in addition, they have severe weight penalties, making them unsuitable. Hence, they will not be further addressed in this review.

Hydrated salts consist of an alloy of inorganic salts and water and enable a cost-effective PCM due to easy availability and low cost. The phase change transformation involves hydration or dehydration of the salts in a process that resembles typical melting and freezing. The salt hydrate may either melt to a salt hydrate containing less water or to an anhydrous form where salt and water are completely separated (Sharma et al., 2009).

Desirable properties of salt hydrates used as PCMs include high latent heat of fusion per unit volume, higher thermal conductivity than organic PCMs and small volume change and easy availability at a lower cost.

Hydrated salts suited for commercial use suffer from incongruent melting. This is a typical problem for hydrated salts as the melting process in many cases leads to the salt releasing water and turning into a salt in its anhydrous from, or a different salt. The density of the anhydrous salt may be higher, causing it to sink to the bottom of the container. When the temperature reaches the freezing point, the salt will be stacked at the bottom and some salt will be unable to reabsorb the water. Hence, the total volume of salt that can undergo phase change has decreased, and the effectiveness of the PCM is reduced. Another issue for hydrated salts is supercooling, which occurs because of poor nucleating properties in many salt hydrates. When supercooling occurs, the liquid phase of a PCM does not freeze (solidify) at the intended temperature, but reaches temperatures lower than the freezing point before actually solidifying. This is a critical flaw as it alters the temperature of the phase change.

3.2.2.3 Eutectic Mixtures

An eutectic is a melting composition of minimum two or more components, each of which melts and freezes congruently. During the crystallization phase, a mixture of the components is formed, hence acting as a single component. The components freeze to an intimate mixture of crystals and melt simultaneously without separation (Lane, 1989). Eutectics can be mixtures of organic and/or inorganic compounds. Hence, eutectics can be made as either organic–organic, inorganic–inorganic or organic–inorganic mixtures (Baetens et al., 2010b). This gives room for a wide variety of combinations that can be tailored for specific applications.

Of organic eutectic mixtures, the most commonly tested consist of fatty acids. Some organic eutectics that have been studied include capric acid/myristic acid (Karaipekli and Sari, 2008), lauric acid/stearic acid, myristic acid/palmitic acid and palmitic acid/stearic acid (Sari et al., 2004), and capric acid/lauric acid (Shilei et al., 2006b). The most common inorganic eutectics that have been investigated consist of different salt hydrates.

The benefits of eutectic mixtures are their ability to obtain more desired properties such as a specific melting point or a higher heat storage capacity per unit volume.

Though it has been given significant interest by researchers, the use of eutectic PCMs for use in (LHTS) systems is not as established as pure compound PCMs. Hence, thermophysical properties of eutectics is still a field for further investigations as many combinations have yet to be tested and proved.

3.2.2.4 Comparison Summary

The advantages and drawbacks for organic, inorganic, and eutectic PCMs are compared and summarized in Table 3.1.

Table 3.1

Overview of advantages and drawbacks for PCMs

Organic Inorganic Eutectics
Advantages Drawbacks Advantages Drawbacks Advantages Drawbacks

ent No supercooling

ent No phase segregation

ent Low vapor pressure

ent Large temperature range

ent Self-nucleating

ent Compatible with conventional construction materials

ent Chemically stable

ent Recyclable

ent High heat of fusion

ent Flammable

ent Low thermal conductivity

ent Low volumetric latent heat storage capacity

ent High volumetric latent heat storage capacity

ent Higher thermal conductivity than organic PCMs

ent Low cost

ent Non-flammable

ent Sharp phase change

ent Corrosive to metals

ent Supercooling

ent Phase segregation

ent Congruent melting

ent High volume change

ent Sharp melting points

ent Properties can be tailored to match specific requirements

ent Limited data on thermophysical properties for many combinations

ent High cost

Image

3.2.3 Encapsulation

As most PCMs designed for building applications go through a liquid phase, encapsulation is needed to avoid problems such as leaking of PCMs to the surface and diffusing of low viscous liquids throughout the material (Özonur et al., 2006). Hence, methods such as direct incorporation and immersion of PCMs in building materials are not well suited for long-term applications (Zhou et al., 2012). As direct incorporation and immersion are techniques that are not widespread for PCMs intended for building applications, they will not be discussed further in this study. PCMs undergo a change in phase; thus it is important to note that this may also lead to change in both volume and thermal conductivity. A large volume change is not desirable as it may put pressure on the encapsulating material.

The material used to encapsulate should not react with the PCM or show signs of deterioration over time. Currently, two main methods are used for encapsulating PCMs, i.e., micro- and macroencapsulation. These two methods give various sizes and shapes of the PCMs and affect how PCMs may be incorporated into a material or construction.

3.2.3.1 Microencapsulation

Microencapsulation of PCMs involves packing the PCM materials in capsules that range from less than 1 µm and up to around 300 µm (Hawlader et al., 2003). The end product consists of the outer shell and the PCM fill in the center. The process can either be performed physically through, e.g., spray-drying (Hawlader et al., 2003) or a coating process (Kaygusuz et al., 2008), or chemically through, e.g., coacervation (Hawlader et al., 2003; Saihi et al., 2006a) or interfacial polymerization (Cho et al., 2002; Liang et al., 2009; Saihi et al., 2006b).

Morphologies of different microencapsulated PCMs are illustrated in Fig. 3.3, which depicts mononuclear, polynuclear, matrix, and multiwall or multifilm microcapsules (Jurkowska and Szczygieł, 2016). The shell materials commonly used for microencapsulation consist of organic polymers or silica. Recently, materials such as SiO2, AlOOH, and TiO2 have been studied (Cao et al., 2015).

image
Figure 3.3 Morphologies of different microencapsulated PCMs, i.e., mononuclear (top left), polynuclear (top right), matrix (bottom left), and multiwall or multifilm (bottom right) (Jurkowska and Szczygieł, 2016).

The advantages of microencapsulation include an increased heat exchange surface giving an increased heat transfer, reduced reactivity with surrounding materials, and the fact that PCMs’ volume can expand without affecting the structure around it. Due to the microencapsulated PCMs’ small size they can be produced in either powder form or dispersed in a liquid (Fig. 3.4) and they do not need additional protection from destruction (Tyagi et al., 2011). This makes it possible to add the PCM directly to materials such as concrete and gypsum without the risk of leakage. However, considerations have to be made as adding microcapsules may affect the structural strength of a material. See also the studies by Zhang et al. (2012) and Zhao and Zhang (2011).

image
Figure 3.4 Microencapsulated PCM dispersed in liquid and as powder (BASF, 2013).

Microencapsulation has shown to be the most promising method of encapsulation for integration of PCMs in building materials. However, microencapsulation leads to a lower latent heat storage capacity per unit volume and unit weight than the pure PCM due to the addition of the encapsulating material.

3.2.3.2 Macroencapsulation

Macroencapsulation refers to PCMs encapsulated in any type of container such as tubes, spheres, or panels, which can be incorporated into building materials or serve as heat exchangers by themselves (Fig. 3.5). The size of these containers is usually larger than 1 cm (Cabeza et al., 2011). Because many PCMs have low thermal conductivity, a disadvantage of macroencapsulation is the tendency to solidify/melt at the edges leading to a slower energy uptake and release, which may prevent the system from discharging completely overnight. The size of the macrocapsules imply that they have to be protected against destruction or perforation and may also make them more difficult to integrate into the building structure, which may make these systems more expensive (Schossig et al., 2005).

image
Figure 3.5 Examples of macroencapsulation of commercial PCMs (RGEES, 2014a,b; Salca, 2014).

3.2.4 Long-Term Stability

For successful large scale application of PCMs into the building sector it is crucial that the PCM and PCM-container system can withstand cycling over an extended period of time without showing signs of degradation. There are two main factors that govern the long-term stability of PCM storage materials: (1) poor stability of the materials, e.g., supercooling and phase segregation; and (2) corrosion between the PCM and the container system (Mehling et al., 2007; Shukla et al., 2008). Degradation of PCMs may result in decreased ability to store latent heat and a difference in phase change temperature.

Accelerated aging tests on stearic acid and paraffin wax, both organic PCMs, have been conducted by Sharma et al. (2002). Both stearic acid and paraffin wax performed well and showed no regular degradation of their melting point over 1500 thermal cycles. However, of the fatty acids, palmitic acid and myristic acid showed the best long-term stability (Sari and Kaygusuz, 2003), which may make them more suited for building applications compared with other fatty acids.

A comprehensive review on the thermal stability of organic, inorganic, and eutectic PCMs has recently been given by Rathod and Banerjee (2013). This work covers the investigations on thermal stability of PCMs done over the past few decades. Paraffins have shown good thermal stability. For fatty acids, the purity plays an important role. Industrial grade fatty acids may experience changes in their thermal behavior over time and should be tested by accelerated aging. Of inorganic PCMs, salt hydrates are the most widely studied. Most studies have shown that the thermal stability of salt hydrates is poor due to phase separation and supercooling. However, the thermal stability may be improved to a certain extent by introducing gelled or thickened mixtures and suitable nucleating materials. In general, new building materials, components and structures should be examined by accelerated climate aging (Jelle, 2012b), PCMs being no exception. Furthermore, a robustness assessment may also be performed (Jelle et al., 2014b).

3.3 State-of-the-Art Phase Change Materials

As mentioned earlier, PCMs can be found in a wide variety of temperature ranges. The PCMs in this study have been limited to PCMs with phase change temperatures in the appropriate range to be efficient in buildings. Cabeza et al. (2011) have listed several tables of PCM properties where the potential areas of use have been divided by the PCMs’ phase change temperature. For use in buildings, three temperature ranges were suggested: (1) up to 21°C for cooling applications, (2) 22–28°C for human comfort applications, and (3) 29–60°C for hot water applications. For this study, PCMs with phase change temperatures ranging from 15 to 32°C have been included. Note that many of the manufacturers included in this study offer PCMs outside this temperature range as well.

3.3.1 Phase Change Material Compounds

By PCM compounds, we refer to products that are made up of PCMs and their prospective encapsulation materials. These are products that have not been combined into products ready for building applications.

Table 3.2 shows an overview of manufacturers and their commercial products with some important properties.

Table 3.2

Manufacturers and properties of commercial PCMs. See the study by Kalnæs and Jelle (2015) for further information

Manufacturer Product Phase change temperature melting/freezing (°C) Latent heat capacity (kJ/kg) Material Typea Encapsulation
BASF DS 5030 21/n.a. 37 Dispersed wax mixture Organic Microencapsulation
DS 5007 23/n.a. 41 Dispersed wax mixture Organic Microencapsulation
DS 5000 26/n.a. 45 Dispersed wax mixture Organic Microencapsulation
DS 5029 21/n.a. 90 Powder form wax mixture Organic Microencapsulation
DS 5008 23/n.a. 100 Powder form wax mixture Organic Microencapsulation
DS 5001 26/n.a. 110 Powder form wax mixture Organic Microencapsulation
RGEES PCM-OM21P n.a./21 120 Organic chemicals Organic  
PCM-HS22P 23/22 185 Inorganic salts Inorganic  
PCM-HS24P 24/25 185 Inorganic salts Inorganic  
PCM-HS29P 29/29 190 CaCl2 and other salts Inorganic  
PCM-OM32P n.a./32 235 Organic chemicals Organic  
Phase change products PC14 14/n.a. 145 Hydrated calcium chloride and calcium bromide Inorganic  
PC17 17/n.a. 145 Hydrated calcium chloride and calcium bromide Inorganic  
PC25 25/n.a. 150 Hydrated calcium and magnesium chlorides Inorganic  
PC29 29/n.a. 188 Hydrated calcium chloride Inorganic  
Entropy Solutions, Inc. PureTemp 15 15/n.a. 165 Vegetable based  Micro- and macroencapsulation
PureTemp 18 18/n.a. 189 Vegetable based  Micro- and macroencapsulation
PureTemp 20 20/n.a. 180 Vegetable based  Micro- and macroencapsulation
PureTemp 23 23/n.a. 203 Vegetable based  Micro- and macroencapsulation
PureTemp 24 24/n.a. 185 Vegetable based  Micro- and macroencapsulation
PureTemp 25 25/n.a. 185 Vegetable based  Micro- and macroencapsulation
PureTemp 27 27/n.a. 200 Vegetable based  Micro- and macroencapsulation
PureTemp 28 28/n.a. 205 Vegetable based  Micro- and macroencapsulation
PureTemp 29 29/n.a. 189 Vegetable based  Micro- and macroencapsulation
PureTemp 31 31/n.a.  Vegetable based  Micro- and macroencapsulation
Salca Thermusol HD26 26/n.a.  Salt hydrate Inorganic Microencapsulation
Thermusol HD32 32/n.a. 150 Salt hydrate Inorganic Microencapsulation
Climator Sweden AB ClimSel C21 21/n.a.  Sodium sulfate Inorganic  
ClimSel C24 24/n.a.  Sodium sulfate Inorganic  
ClimSel C28 28/n.a.  Sodium sulfate Inorganic  
ClimSel C32 32/n.a.  Sodium sulfate Inorganic  
Phase Change Material Products Ltd. S15 15 160 Salt hydrate Inorganic  
S17 17 160 Salt hydrate Inorganic  
S19 19 160 Salt hydrate Inorganic  
S21 22 170 Salt hydrate Inorganic  
S23 23 175 Salt hydrate Inorganic  
S25 25 180 Salt hydrate Inorganic  
S27 27 183 Salt hydrate Inorganic  
S30 30 190 Salt hydrate Inorganic  
S32 32 200 Salt hydrate Inorganic  
A15 15 130  Organic  
A16 16 213  Organic  
A17 17 150  Organic  
A22 22 145  Organic  
A22H 22 216  Organic  
A23 23 145  Organic  
A24 24 145  Organic  
A25 25 150  Organic  
A25H 25 226  Organic  
A26 26 150  Organic  
A28 28 155  Organic  
A29 29 226  Organic  
A32 32 130  Organic  
X25 25 110   Solid–solid phase change
X30 30 105   Solid–solid phase change
PCM Energy P. Ltd. Latest 18T 17–19 175 Inorganic salts Inorganic  
Latest 20T 19–20 175 Inorganic salts Inorganic  
Latest 25T 24–26 175 Inorganic salts Inorganic  
Latest 29T 28–30 175 Inorganic salts Inorganic  
Latest 32S 31–32 >200 Inorganic salts Inorganic  
Microtek Laboratories MPCM 18 18/n.a. 163–173 n-Hexadecane Organic Microencapsulation
MPCM 18D 18/n.a. 163–173 n-Hexadecane Organic Microencapsulation
MPCM 24 24/n.a.  Special blend  Microencapsulation
MPCM 24D 24/n.a.  Special blend  Microencapsulation
MPCM 28 28/n.a. 180–195 n-Octadecane Organic Microencapsulation
MPCM 28D 28/n.a. 180–195 n-Octadecane Organic Microencapsulation
MPCM28D-IR 25–32/n.a. 160–180 Special blend  Microencapsulation
MPCM 32 32/n.a.  Special blend  Microencapsulation
MPCM 32D 32/n.a.  Special blend  Microencapsulation
Cristopia AC 27 27    Nodules filled with PCM
Rubitherm Technologies GmbH RT 18 HC 17–19/19–17 250    
RT 21 18–23/22–19 160    
RT 21 HC 20–23/21–19 190    
RT 22 HC 20–23/23–20 200    
RT 24 21–25/25–21 150    
RT 25 22–26/26–22 148    
RT 25 HC 22–26/26–22 230    
RT 27 25–28/28-25 179    
RT 28 HC 27–29/29–27 245    
RT 31 27–33/33–27 170    
SP 21 E 22–23/21–19 160 Salt hydrates and organic compounds   
SP 22 E 22–23 180 Salt hydrates and organic compounds   
SP 24 E 24–25/23–21 222 Salt hydrates and organic compounds   
SP 25 E 24–26/24–23 200 Salt hydrates and organic compounds   
SP 26 E 25–27/25–24 200 Salt hydrates and organic compounds   
SP 31 31–33/30–28 220 Salt hydrates and organic compounds   
PX 15 10–17/17–10 85   Microencapsulation
PX 25 22–25/25–22 96   Microencapsulation
PX 27 25–28/28–25 102   Microencapsulation
PX 31 27–31/33–27 110   Microencapsulation

ImageImage

aMany manufacturers do not give a full description of the mixtures used in PCMs. Hence, type has only been limited to organic and inorganic, though some of these may be eutectic mixtures as well.

3.3.2 Phase Change Materials in Products for Building Applications

PCM technology has already been combined with several other products that are directly aimed at the building sector. PCM-enhanced materials include, e.g., wallboards, floor tiles, ventilation systems, mats, and tubes. Table 3.3 shows an overview of manufacturers and their commercial products with some important properties.

Table 3.3

Manufacturers of products ready for building applications. See the study by Kalnæs and Jelle (2015) for further information

Manufacturer Product Phase change temperature (°C) Latent heat capacity (kJ/kg) PCM material Typea Building product
Dupont Energain 18–24 515 kJ/m2 Paraffin wax compound Organic Thermal mass panels
Knauf Comfort board   BASF micronal  Gypsum board with microcapsules of PCM
Smartboard 23 23 200 kJ/m2 BASF micronal   
Smartboard 26 26 330 kJ/m2 BASF micronal   
RGEES PB29P 29 33 kWh/cbm  Organic Spheres of 75 mm in diameter filled with PCM
PB22P 22 0.1 kWh/(ft2)  Inorganic Tubes filled with PCM
PB24P 24 0.1 kWh/(ft2)  Inorganic Tubes filled with PCM
PB29P 29 0.1 kWh/(ft2)  Inorganic Tubes filled with PCM
Phase Change Energy Solutions BioPCMat M27 23  Bio-based Organic Mats filled with PCM
BioPCMat M51 25  Bio-based Organic Mats filled with PCM
BioPCMat M91 27  Bio-based Organic Mats filled with PCM
Thermastix   Bio-based Organic Stick filled with PCM
Dörken Delta-cool 24     No longer available on the market?
Delta-cool 28     No longer available on the market?
Salca K-Block  590 kJ/m2 Salt hydrate Inorganic Mat filled with PCM
SGL Group Ecophit GC20 22 85 kJ/kg    
Ecophit LC20 22 140 kJ/kg    
Phase Change Material Products Ltd. FlatICE     Flat container filled with PCM
TubeICE     Tubes filled with PCM
BallICE     Spheres filled with PCM
National Gypsum ThermalCORE   BASF micronal  Gypsum board with microcapsules of PCM covered with a fiberglass mat
H + H Deutschland CelBLoc Plus   BASF micronal  Concrete blocks with microcapsules of PCM
Maxit Deutschland Maxit clima     Plaster with integrated microcapsules of PCM
Ilkazell Isoliertechnik Ilkatherm air conditioning systems      
Monodraught Coolphase     Ventilation system with PCM
Tate EcoCore 75.2°F 147 Btu Vegetable bio-based Organic Floor tiles
Emco Emcovent     Decentralized PCM modules
Autarkis      PCM heat exchangers
Armstrong World Industries Coolzone   BASF micronal Organic Chilled roof system
Trox FSL-B-PCM     Supply air unit with PCM

Image

aMany manufacturers do not give a full description of the mixtures used in PCMs. Hence, type has only been limited to organic and inorganic, though some of these may be eutectic mixtures as well.

3.3.3 Phase Change Materials in Windows

Windows incorporating PCMs are highly limited on the market. There can be several reasons for this. First of all, the PCM relies on a process where the materials solidify below a certain temperature. Windows are in many cases used for aesthetic purposes, to allow visibility towards the outside or allow daylight in to the building. Most PCMs in use on the market are not transparent in both their liquid and solid state. Hence, PCM windows will be blurry and reduce transmission of daylight and solar radiation in general. Only one manufacturer of windows including PCMs has been found in this study. The products currently on the market are categorized as translucent. Table 3.4 shows an overview of the manufacturers of PCM windows and some of their important properties.

Table 3.4

Manufacturers of windows incorporating PCMs. See the study by Kalnæs and Jelle (2015) for further information

Manufacturer Product Phase change temperature (°C) Heat storage capacity (Wh/m2) Tvis solid/liquid (%) Material Type
GLASSX GLASSX crystal 26–30 1185 0–28/4–45 Salt hydrates Inorganic
GLASSX comfort 26–30 1185 0–38/4–55 Salt hydrates Inorganic
GLASSX slim 26–30 1185 0–38/4–45 Salt hydrates Inorganic
GLASSX store 26–30 1185 0–38/4–55 Salt hydrates Inorganic

Image

3.3.4 Comparison of Commercial Products

Commercial PCMs come in a variety of forms and solutions, e.g., PCM-infused wallboards, microcapsules filled with a PCM dispersed in a liquid or in powder form or mats that can be installed directly. There is no clear preference towards organic or inorganic PCMs in commercial products and both are ready and available for use. A market study performed by Markets and Markets (2013) showed that organic PCMs held the highest market share in terms of value due to higher cost, while inorganic PCMs held the highest market share in terms of volume.

The tables include manufacturers that produce the raw PCM material, as well as manufacturers that produce building materials based on their own or other manufacturers’ PCMs. All commercial building materials produced with PCMs use some form of encapsulation. None of the commercial building products found in this study are created using direct incorporation or immersion. This may be due to the fact that these techniques will lead to a material with a large probability of leakage of the PCMs to the surface. Hence, the material does not meet the strict requirements for building applications with regard to lifetime expectancy.

Both products created by macro- and microencapsulation are already commercialized. Macroencapsulated products include tubes, flat containers, and spheres ranging from a few millimeters up to about 75 mm. Microencapsulated PCMs are generally already integrated into building materials such as gypsum and plaster.

Several producers state that there is no sign of degradation after 10 000 cycles for their organic PCMs, giving the PCMs a lifetime of approximately 30 years or more.

An important aspect is how the various manufacturers have solved the issue of organic PCMs with respect to fire resistance. Materials that are aimed for building applications have to overcome some strenuous requirements with regard to safety. And as mentioned earlier, one drawback of organic materials is the fact that many of them are flammable and may release toxic fumes during combustion.

An interesting point is the phase change temperature. Some manufacturers state both melting and freezing temperatures, whereas some state a “phase change temperature”, which does not tell whether the material freezes or melts at the same temperature or if it is simply one of the two temperatures, and finally some manufacturers state the range over which freezing and melting occur instead of a single temperature where the phase change is at its peak. The variation in how manufacturers choose to state the phase change temperature is also an interesting point with respect to the uncertainty of describing the enthalpy curve of the PCMs that still exists, and the difficulty of giving a value that properly represents the phase change reaction.

The latent heat capacity also varies for the presented products, and it is hard to determine why there are such differences as the materials used are not always given. The parameters that affect a PCM’s latent heat capacity should be more clearly stated, to enable comparison of products. The total latent heat capacity does after all describe the total energy that can be absorbed during phase change, and is one of the most vital parameters when selecting PCMs. By evaluating the studied products the latent heat capacity seems to vary due to encapsulation methods and the application method of the PCM, e.g., slurries or powder. Altering of the phase change temperature of the same product from the same manufacturer also yields changes to the latent heat capacity; this is most likely due to changes in the PCM mixture to attain the desirable temperature. What changes are made in the material to attain a desired phase change temperature is not so clearly defined, so it has not been possible to determine which parameters have been altered and why this affects the latent heat capacity.

3.4 Phase Change Materials in Building Applications

3.4.1 Building Applications

Zhu et al. (2009) presented an overview of research conducted on PCMs with regard to their dynamic characteristics and energy performance in buildings. This overview divided the possible building applications of PCMs into four categories:

1. free cooling

2. peak load shifting

3. active building systems

4. passive building systems.

Besides these suggested applications, another interesting possible benefit of PCMs is their ability to increase thermal comfort by smoothing out temperature fluctuations. The study given herein is mostly focused on the effect of passive building systems through the integration of PCM-enhanced building materials. However, a short overview and explanation of some of the research that has been performed and possibilities for the other systems will be given.

3.4.1.1 Free Cooling

Free cooling systems with PCMs work by storing outdoor coolness (e.g., during the night) and release the coolness indoors during the day. The PCM can then be used during the day to absorb the heat from, e.g., passing air in a ventilation system or water in a pipe system, and stored as latent heat, to cool the building in the day when temperatures are higher and the need for cooling arises. These systems work as long as the ambient temperature allows the PCM to freeze and melt over the day, i.e., the ambient temperature must be above the phase change temperature during the day and below it during the night (Fig. 3.6) (Zalba et al., 2004).

image
Figure 3.6 Schematic of a free cooling air conditioning system according to Zalba et al. (2004) (top) and Mosaffa et al. (2013) (bottom).

Mosaffa et al. (2013) described a free cooling system using PCM slabs separated by air gaps for air to flow through (Fig. 3.6). The model developed from this can be used to evaluate performances of latent heat storage systems for free cooling air-conditioning systems.

3.4.1.2 Peak Load Shifting

Peak loads that hit during the day put pressure on the electrical grid and also lead to the need for heating, ventilation, and air conditioning (HVAC) systems being dimensioned for higher heating or cooling loads. Ultimately, this could lead to a need for more power generation facilities being built. By shifting the peak load away from the peak hours of electrical demand using PCMs, the peak load may be divided throughout the day, reducing the highest peaks (Halford and Boehm, 2007). Fig. 3.7 illustrates how the peak may be both reduced and shifted by the use of PCMs.

image
Figure 3.7 Illustration of peak load offset and peak load reduction. Redrawn from Tate (2013).

Sun et al. (2013) reviewed strategies involving PCMs for peak load shifting and control that have been tested so far. From the studies reviewed, peak cooling load reductions ranging from 10% to 57% with no or simple control strategies were found. The greatest reductions were found in cases where the PCM was compared against an insulated lightweight construction, while the lowest reduction was found when comparing against structures containing more mass, e.g., concrete. However, the cost saving potential of these systems could be further improved if more sophisticated load shifting control strategies were developed.

3.4.1.3 Active Building Systems

The storage capability of PCMs can be integrated into systems such as, e.g., solar heat pump systems, heat recovery systems and floor heating systems. Such systems can be combined to attain a peak load reduction as described in the previous chapter. However, if they are made even more effective they can achieve further savings through reduced electrical demand for HVAC systems. As an example, see the study by Fiorentini et al. (2015).

An example of a radiant floor incorporating PCMs in an active system has been described by Ansuini et al. (2011). The system investigated consists of a lightweight piped radiant floor system with an integrated PCM layer aimed at buffering internal gains during the summer season without affecting the winter warming capacity.

3.4.1.4 Passive Building Systems

Passive building systems and their use have attained the most interest so far. For passive applications, PCMs are integrated into the building envelopes to increase the thermal mass. This is especially beneficial in lightweight constructions, which suffer from low thermal inertia. A known issue for these buildings is large temperature fluctuation in the summer due to excessive overheating caused by a lack of thermal mass. This is especially the case in cold climates, where buildings have been built according to passive house standards, often involving large amounts of thermal insulation to reduce heating loads in the winter.

The materials incorporating PCMs will melt during the daytime and solidify during nighttime. This will help prevent rooms from overheating during the daytime in warm months and may also reduce the need for heating during the nighttime in the winter.

An issue that has been brought up is the importance of getting passive PCM systems to completely discharge during the nighttime in warm periods. If the PCM is not able to completely solidify, the effectiveness of the system may be considerably reduced. This point makes PCMs more effective in climates with large daily variation in temperatures. For areas where the discharge does not happen naturally, cool air has to be supplied during the nighttime to reset the PCMs completely.

3.4.1.5 Thermal Comfort Control

Though PCMs show potential for energy savings, another important factor to highlight is the benefits PCMs may have towards increasing the overall indoor thermal comfort. Lan et al. (2011) showed a correlation between workers’ performance and productivity compared to the sensation of thermal comfort due to shifting temperatures. Seppänen and Fisk (2006) showed that elevated air temperatures had a negative effect on performance and productivity. When temperatures increased up to 20°C there was an increase in working performance. However, when temperatures increased above 23°C there was a decrease in productivity. Maintaining a steady temperature around the comfort zone for longer periods without relying on HVAC systems may be possible with PCMs.

With PCMs installed, temperature fluctuations are reduced. The focus should be placed on selecting a PCM within the desired melting/freezing point so temperatures stay stable around the comfort temperature. This will benefit the indoor climate in two ways. First, the temperature will be held more stable, reducing the feelings of thermal discomfort due to temperature fluctuations throughout the day. Second, the peak temperature will be reduced and should not reach a temperature that leads to increased thermal discomfort. Another possible benefit of PCMs may be that they lead to a more uniform temperature between surfaces and air temperature, reducing thermal discomfort from radiative heat.

3.4.2 Solar Energy Storage

PCMs hold the ability to store energy given off by the sun. Where solar cell panels can produce energy during hours of solar radiation, PCMs can store some of the excess energy and release it at a more needed time of the day. This can be combined with different energy distribution systems such as a heat pump.

For PCMs to best utilize the solar energy it is important that the surfaces are positioned in areas where they can fully absorb the energy coming from the sun, e.g., wallboards or PCM windows. PCM incorporated into concrete floors is an example of a position that poorly utilizes the possible energy gains from the sun, as such floors are covered with other materials in most traditional structures.

3.4.3 Examples of Integration of Phase Change Materials for Passive Systems

Though there have been few detailed studies on the overall effect of PCMs in real life constructions, commercial PCM products have already been used in several projects. In the following, examples from some of these constructions, reasoning for the use of PCMs, as well as general thoughts about possible areas and materials where PCMs can be incorporated beneficially, will be explored. The solutions treated in this chapter are all seen as part of a passive system.

When used in buildings, PCMs can be integrated into other building materials. This has attracted a lot of interest as it will enable buildings to be built fairly similar to the way they are built today, but with materials that have an increased thermal energy storage capacity.

In the reviewed literature it is obvious that wall systems integrating PCMs have received the most attention. Studies on floor and roof systems are limited.

3.4.3.1 Walls

The most common solution for implementing PCMs into buildings so far is by installing PCM enhanced wallboards towards the interior side of the building envelope. For lightweight structures, which have low thermal inertia, PCMs can provide a significant increase in thermal storage capacity. When facing the interior rooms of a building or being used in partition walls, wallboards will be able to absorb and release heat throughout the room for large parts of the day. Several investigations, both experimental and practical, have been conducted to see how this affects overall indoor climate and energy usage.

Experimental hot box tests have shown that PCM wallboards give an obvious reduction in temperature fluctuations, as demonstrated in Fig. 3.8 (Cao et al., 2010). The dynamics of external wall structures with PCMs investigating the possibility of substituting thick and heavy thermal mass external walls used in high latitude countries by thin and light thermal mass ones were studied by Chwieduk (2013).

image
Figure 3.8 Indoor air and surface temperatures of a test wall with and without PCM panels versus elapsed time, demonstrating the ability of the PCM panels to decrease the indoor temperature variations with about 2°C. Redrawn from Cao et al. (2010).

A full-scale investigation was performed on a lightweight building’s inner partition walls. In this study, several positive effects could be seen when a room with PCM wallboards was compared to one without PCMs added. Air temperature fluctuations were reduced, the overheating effect was lower, and the PCM wall was shown to release energy when temperatures fell. The tests also showed reduced fluctuations of the surface temperature of the wall, i.e., lower surface temperatures during the day and higher surface temperatures during the night. Hence, the thermal comfort by radiative effects could be increased during daytime (Kuznik et al., 2008).

Kuznik and Virgone (2009) performed an experimental investigation in a full-scale test room for a typical day in winter, summer, and midseason with PCM enhanced wallboards. The PCM wallboards were created by adding 60% microencapsulated paraffins in a flexible sheet of 5 mm thickness. The PCMs were shown to reduce overheating effects, reduce the surface temperature on the walls, and enhance the natural convection mixing of the air.

Evola et al. (2013) performed a simulated case study of an office building refurbished with PCM enhanced wallboards during summer conditions. The wallboards were made of an aluminum honeycomb matrix and filled with 60% microencapsulated PCMs with paraffin as the core material. This simulation showed that even if the PCMs are frequently activated, on average they only utilized 45% of their total latent heat storage potential, meaning that the entirety of the PCM will not melt or solidify within each cycle. The utilization of total latent heat storage potential is affected by factors such as, e.g., convective heat transfer across the wall surface, whether the placement is in an area of a room that receives low amounts of direct solar radiation, and climate conditions such as cloudy weather or extreme temperatures that are outside the expected normal temperatures. The study also discussed the importance of evaluating the PCM over a longer period of time, such as a few months rather than a few days, to be able to better evaluate the activation rate and the utilization rate of the total latent heat storage potential of a PCM under a given climate.

Diaconu (2011) studied the influence of the occupancy pattern and ventilation that may be affected by PCM enhanced envelopes. Through numerical simulations the potential for thermal energy savings for heating was investigated. It was clear that occupancy patterns and ventilation must be considered when attempting to select the optimal PCM melting temperature.

Ascione et al. (2014) investigated the possibility of refurbishing a building with PCM plaster on the inner side of the exterior building envelope and the effect it would have on energy savings and indoor comfort in the cooling season. The results were simulated while varying the phase change temperature, thickness of the wallboard and the location of the PCM layer. Phase change temperatures ranging from 26°C to 29°C were tested. With a melting temperature of 29°C the highest energy saving potential and increase in comfort hours were seen. However, the achievable benefit for energy savings in climates simulating Seville and Naples were no more than 3%, while Marseille and Athens received a benefit of 4.1% and 3.5%, respectively. The highest energy saving effect was reached in Ankara, with energy savings of 7.2%. The comfort hours during the occupied hours increased by 15.5% (Seville), 22.9% (Naples), 19.8% (Marseille), 15.8% (Athens), and 20.6% (Ankara).

The experiments conducted by Ascione et al. (2014) also highlight another important factor for PCMs. They found that during the summer, the period with temperatures lower than the phase change temperature, i.e., the period where the PCM solidifies, would not always be sufficient for the PCM to fully solidify. The same would also occur during the winter, when the heat available during the day would not be enough to completely melt the PCM. This indicates that the optimal phase change temperature is seasonal, and complete discharge, or complete absorbance of heat, is difficult to obtain for a single PCM.

Shilei et al. (2006a) tested the impact of PCM incorporated gypsum boards in winter climate conditions in the northeast of China. The experiment was performed on two similar rooms, one with PCMs and one without PCMs. The PCM room showed that it could reduce the heat transfer to outdoor air and improve the thermal comfort. Furthermore, the possibility to reduce the scale of the heating equipment was shown.

A renovation project in the south of Lyon using PCM enhanced wallboards was monitored over a 1-year period by Kuznik et al. (2011b). The room was compared to a room in the same building that was renovated without PCMs. The PCMs showed increased thermal comfort of the occupants due to air temperature and radiative effects. However, the PCMs seemed to be unable to utilize their latent heat storage capacity over several periods over the year due to excessive temperatures and the lack of a complete discharge overnight.

Becker (2014) investigated how thermal and energy performance of a building would be affected by placing PCMs on the inner surface of the walls during summer conditions. By using EnergyPlus and a simplified model, the thickness of the PCMs needed to store heat gains from solar radiation and occupancy patterns occurring during the day, and ventilation needed by night to discharge the system, were studied for lightweight buildings, semilightweight buildings, and heavyweight buildings in a Mediterranean climate. In heavyweight constructions, PCMs could improve thermal conditions, but had a marginal effect on energy savings. For lightweight dwellings and offices and semilightweight schools the analyses showed positive effects on both thermal and energy performance. The greatest effect was seen in lightweight office buildings, where energy savings up to 57% were found. However, this study also pointed out the importance of occupancy patterns and the focus on including night ventilation needed for discharging when studying PCMs. It is still important to get full-scale validation in occupied buildings under normal working and climate conditions.

3.4.3.2 Floors

Areas that are in direct contact with solar radiation hold large potential for storage of thermal heat energy. Floor solutions incorporating PCMs in areas of a building where the sun shines for large parts of the day may benefit from incorporating PCMs. Fig. 3.9 shows flat profiles filled with PCMs that were used in the floor in North House, a competitor in the US Department of Energy’s solar decathlon.

image
Figure 3.9 Flat profiles that can be installed underfloor to store and release latent thermal heat energy (Cosella-Dörken, 2014a).

Xu et al. (2005) performed a simulation of the thermal performance of PCMs used in a passive floor system during the winter season. The performance on the systems was influenced by the choice of covering material, the air gap between the PCM and covering material, and the thickness of the PCM. For the simulations performed, the thickness should not exceed 20 mm as this would not increase the influence of the thermal storage significantly.

3.4.3.3 Roofs

Implementing PCMs into roof systems does not seem to have received much attention. Only a few studies on the possible effects of PCMs in passive roof systems have been found. The thought is that PCMs placed on the roof will be able to absorb the incoming solar energy and the thermal energy from the surroundings to reduce temperature fluctuations on the inside.

Pasupathy and Velraj (2008) studied the effects of a double layer of PCM for year round thermal management in Chennai, India. An experiment was performed with a PCM roof panel compared to a reference room without the PCM panel. The PCM used was an inorganic eutectic of hydrated salts. The experiment showed that the PCM panel on the roof narrowed the indoor air temperature swings, and that such a system could perform during all seasons when the top panel had a melting temperature 6–7°C higher than the ambient temperature in the early morning during the peak summer month, and the bottom panel had a melting temperature near the suggested indoor temperature.

Kosny et al. (2012b) set up a naturally ventilated roof with a photovoltaic (PV) module and PCMs to work as a heat sink (Fig. 3.10). The goal was that the PCM would absorb heat during the day in winter and release it in the night to reduce heating loads. In the summer the PCM would absorb heat to reduce the cooling loads in the attic beneath. A full scale experiment was performed over a whole year from November 2009 until October 2010 in Oak Ridge, Tennessee, USA. The data from the tests were compared with a conventional asphalt shingle roof. The PV-PCM attic showed a 30% reduction in heating loads during the winter and a 55% reduction in cooling loads. Furthermore, a 90% reduction in peak daytime roof heat fluxes was observed.

image
Figure 3.10 PV-PCM roof (Kosny et al. 2012b).

3.4.3.4 Windows and Shutters

Windows represent a part of the building that is considered to lead to a higher energy consumption. In warm climates dominated by cooling loads, excessive solar heat gains lead to an increased need for mechanical cooling. In cold climates, much of the energy escapes through glazed facades, leading to a need for mechanical heating (Ismail et al., 2008). Several new advanced window technologies such as electrochromic windows, low-emissivity glazing, evacuated glazing, self-cleaning glazing, building integrated photovoltaics (BIPV) as solar glazing, etc., have been explored to counter these issues (Baetens et al., 2010d; Jelle, 1993, 2013a, 2013b, 2015; Jelle and Hagen, 1993, 1999a; Jelle and Breivik, 2012a,b; Jelle et al., 1998, 2007, 2012a,b, 2015b; Midtdal and Jelle, 2013). However, glazed facades still suffer from low thermal inertia, and have no way of storing excess heat. Transparent PCMs for use in windows represent an opportunity that has been explored for this purpose.

The first issue to overcome is the desire to have sufficiently transparent windows. Fig. 3.11 shows a commercialized glazing filled with a PCM, also depicting the visibility in its liquid state. As of today, only translucent PCMs have been used for PCM windows; though they enable relatively high amounts of visible light to pass through, they do not offer the same visibility as regular windows.

image
Figure 3.11 Illustration of a PCM filled window (left) and commercialized PCM window in its liquid state (right) (GlassX, 2015).

The solar transmittance of a commercial grade PCM was tested by Jain and Sharma (2009). For a pure PCM with a thickness of 4–30 mm the solar transmittance was found to be 90.7–80.3%, respectively. Due to the fact that PCMs have low thermal conductivity, they concluded that PCMs may be interesting as a transparent thermally insulating medium.

Goia et al. (2013) compared a prototype PCM glazing system with a traditional double glazed insulating unit with regard to the effect on thermal comfort in the winter, summer, and midseason. The two glazing systems were measured over a 6-month experimental campaign, and the data was used to numerically evaluate an indoor office room. The thermal conditions were considerably improved for most parts of the year with the PCM prototype compared to the regular double glazing. However, on cloudy days, the two systems showed similar properties on thermal comfort. The study also highlighted the importance of selecting the correct melting temperature for the PCM as this could affect the system negatively if not optimized.

A similar study was performed by Weinläder et al. (2005), comparing a double glazing unit combined with PCM compared to a regular double glazing unit. The test specimens were studied experimentally in an outdoor test facility and the data gained from the experiment was used for further simulations. The light transmittance from this solution was found to be 0.4, giving them the ability to be used as daylighting elements. Compared to regular double glazing they gave a more equalized energy balance, achieving moderate heat gains with very low heat losses. During the winter season, PCM windows improved the thermal comfort considerably and in the summer they shifted heat gains to later in the evening. However, the PCM windows cannot be used where visual contact to the environment is desired.

Goia et al. (2014) performed a full-scale test on a PCM glazing prototype. The test was performed on a south facing wall during summer, midseason, and winter days in a subcontinental climate and compared to a conventional double glazing for reference. The experiment showed that PCM glazing can reduce the energy gain in the summer by more than 50%. In the winter, heat loss reduction during the day was observed; however, this technology might not be effective if the purpose is to achieve solar heat gains. The experiment also pointed out the importance of removing the stored heat during the night via, e.g., night cooling, and selecting the correct phase change temperature, if the system is to function optimally.

Grynning et al. (2013a) performed measurements on a state-of-the-art commercial glazing including a PCM in a large scale climate simulator. The tested glazing was a 4-layer glazing with a prismatic glass in the outer pane and a PCM fill in the cavity between the inner panes. The study showed that characterization of static components (e.g., U-value, solar heat gain coefficient, thermal transmittance) is insufficient for describing the performance of PCMs due to their dynamic nature.

A zero energy office building using translucent PCMs in the window construction has been built in Kempen, Switzerland (Fig. 3.12). In this project, each second window panel has been equipped with PCM windows. The aim for the windows is to effectively store solar energy during the warmer parts of the day, and release thermal energy throughout the colder periods of the day to reduce the total energy required for space heating.

image
Figure 3.12 Zero energy office building from Kempen, Switzerland (Cosella-Dörken, 2014b).

Alawadhi (2012) investigated the possibility to implement PCMs in window shutters, where the goal was to see if the solar heat could be absorbed before it reached the indoor space. When compared to foam shutters, the shutters containing a PCM could lead to a reduction in heat gain through the windows by as much as 23.29%.

3.4.3.5 Concrete

Adding PCMs directly into concrete has shown some promising results through lower thermal conductivity and an increase in thermal mass at specific temperatures. However, PCM concrete has shown some undesirable properties such as lower strength, uncertain long-term stability, and lower fire resistance (Ling and Poon, 2013). Several studies have been conducted on PCM-concrete and have shown positive effects through reduced indoor temperatures in warm climates (Cabeza et al., 2007).

Combining concrete structures with PCMs has been tried in several different ways. One studied solution is to drill holes in the concrete, which may then be filled with a PCM (Alqallaf and Alawadhi, 2013).

Royon et al. (2013) tested the possibility of filling the already hollow areas in a hollow concrete floor with PCMs. The concrete was filled with a paraffin PCM with a melting temperature of 27.5°C. This test showed that the temperature on the other side of the hollow concrete was lower during summer conditions. Hence, such floors can be used as a passive thermal conditioner during the summer. However, more tests are needed with real life climate conditions to validate the effects.

3.4.3.6 Thermal Insulation Materials

Incorporation of PCMs into fibrous thermal insulation materials has received considerable attention. Kosny et al. (2012a) performed an experimental and numerical analysis of a wood-frame wall containing PCM enhanced fiber insulation. The wall assembly had an R-value of 4.14 (m2 K)/W (U-value of 0.241 W/(m2 K)). For fiber insulation filled with 30 wt% PCM in summer conditions, results showed a reduced peak hour heat gain of 23–37% in Marseille and 21–25% in Cairo and Warsaw.

3.4.3.7 Furniture and Indoor Appliances

A point that has not been investigated for this study, but should be mentioned, is the possibility of using PCMs in furniture and other indoor appliances. The benefit of PCMs is as mentioned their ability to store heat in periods where there is a surplus, and release the heat when there is a deficit. It would be interesting to study how incorporation of PCMs into other components in a building besides the structural components could benefit energy savings and thermal comfort. PCMs have already been widely studied for textile applications (Sarier and Onder, 2012), showing that there is a possibility of adding PCMs to various forms of materials.

3.4.4 Retrofitting

Building retrofitting is an important measure to reduce the total energy usage worldwide. PCMs may offer an increase in a building’s overall energy efficiency with little or no additional space required (Rodriguez-Ubinas et al., 2013). The impacts on the design of the building can be minimized through solutions such as PCM enhanced wallboards and PCM shutters. See also the studies by, e.g., Ascione et al. (2014), Bourne and Novoselac (2015), Evola et al. (2013), Fiorentini et al. (2015), Kuznik et al. (2011b) and the various ones described within this text.

3.4.5 Safety Requirements

The safety requirements for materials used in buildings are crucial points for PCMs to fulfill. As mentioned earlier, PCMs should not be toxic or flammable. However, for many organic PCMs flammability and possible release of toxic fumes during combustion have been an issue. Solutions have been made to counter this issue, such as ignition resistant microcapsules for PCMs and the adding of fire retardants.

Hence, it is of significance that manufacturers of PCMs for building applications are required to give reliable information about the fire performance of their products. Nguyen et al. (2013) reviewed the work that has been carried out to improve fire safety of PCMs. The work investigated the use of fire retardants to increase fire resistance of composite PCMs.

3.5 Future Research Opportunities

3.5.1 Improving the Current Technologies

3.5.1.1 Increasing Thermal Storage Capacity

A desirable trait for PCMs is a high thermal storage per unit volume and unit weight. Hence, there is always a focus on finding new materials and solutions that may increase the thermal storage capacity for a given volume or weight of a PCM. However, as mentioned earlier, this should not come at the cost of using potentially environmentally harmful materials and chemicals.

Microencapsulation reduces the risk of PCMs leaking from the material in their liquid state. Unfortunately, the encapsulation leads to a lower latent heat storage capacity. The method is a promising way for integrating PCMs with building materials, so finding new methods or materials to encapsulate with, which will give a better thermal storage capacity, is desirable. Possibilities to achieve this may include developing a thinner shell for the capsules, which will increase the weight percent of PCMs in the final products.

Research with miscellaneous microcapsules will be continued, and which morphologies and types of microencapsulated PCMs, as depicted in Fig. 3.13, and even other configurations not hitherto known or explored, will come out as the better ones, remain to be seen. The exploration of these comprises many parameters including, e.g., material types, dimensions, and interfaces. Microencapsulated PCMs may also be incorporated in various materials; one example with gypsum boards is shown in Fig. 3.14.

image
Figure 3.13 Morphologies and types of microcapsules for microencapsulated PCMs. Irregular shape (A top), simple (B top), multiwall (C top), multicore (D top), matrix particle (E top), mononuclear (A bottom), polynuclear (B bottom), matrix encapsulation (C bottom), and multifilm (D bottom) (Giro-Paloma et al., 2016).
image
Figure 3.14 Scanning electron microscope (SEM) images of micro-PCM (A), enlarged single microcapsule (B), gypsum boards with micro-PCM/gypsum weight ratios 30/70 (C), 50/50 (D), and 60/40 (E) (Zhang et al., 2012).

3.5.1.2 Enhancing Heat Transfer

A problem that has been addressed throughout the literature is the low thermal conductivity for many promising PCMs (around 0.15–0.2 W/(mK) for organic PCMs and around 0.5 W/(mK) for inorganic salts). Low thermal conductivity reduces the rate of heat absorption or heat release throughout the PCM, i.e., reducing the effectiveness at which it can store and release thermal energy. This may lead to a system that does not fully utilize the full latent heat storage of PCM materials. In the study conducted on wallboards under summer conditions by Evola et al. (2013) they found that the wallboards tested only took advantage of about 45% of the possible latent heat storage. Note that this is affected by several other factors than thermal conductivity alone, such as local climate conditions and the way the system is installed.

Fan and Khodadadi (2011) reviewed methods that have been used to enhance the thermal conductivity for the last few decades. Recently, graphite based PCM systems and metal foams have been getting increased attention.

One solution that has been investigated includes adding of a material with a high thermal conductivity to the PCM. A set of materials that has been investigated for this purpose is various carbon-based nanofillers (Babaei et al., 2013; Ji et al., 2012; Yu et al., 2014). Though the results vary, the increase in thermal conductivity has increased between 65% and 336% in these studies.

Another solution is to add the PCM to a material with a porous structure and a high thermal conductivity (Fig. 3.15). Investigations on graphite foam composites (Sedeh and Khodadadi, 2013; Zhong et al., 2010; Song et al., 2012) and metal foams (Jiang et al., 2012; Zhao et al., 2010) have been conducted. Graphite foam composites have shown potential for creating a structure with high thermal conductivity (ranging from 230–570 times higher than the original PCM).

image
Figure 3.15 Illustration of (A) graphite foam, (B) PCM material, and (C) graphite foam and PCM composite (Sedeh and Khodadadi, 2013).

Note that adding a new material to the PCM and creating a composite material reduce the total latent heat storage ability per unit weight due to the adding of new particles. The energy uptake and release of the system may improve, but as of yet, this comes at the cost of less thermal heat storage per unit weight.

3.5.2 New Technologies

3.5.2.1 Nanoencapsulated Phase Change Materials

Though it has not yet been used in commercial products, nanoencapsulation of PCMs may have potential to improve PCM materials. The idea is similar to that of microencapsulation, but with even smaller capsules ranging in the size of nanometers. A smaller size of the capsules could be interesting due to a higher heat transfer area, leading to a higher heat transfer speed. Several studies have been performed and given functional nanocapsules with PCMs.

Chen et al. (2012) prepared nanocapsules containing n-dodecanol by miniemulsion polymerization. These capsules reached a latent heat capacity of 98.8 J/g.

Rao et al. (2013) performed a simulation on nanocapsules by using n-octadecane as the core material and SiO2 as the shell material. This study pointed out that the possibilities of creating excessive thick or thin shells might be problematic for the current nanocapsules.

Latibari et al. (2013) created nanocapsules containing palmitic acid as the core material and SiO2 as the shell material by a sol-gel method. The latent heat was measured by a differential scanning calorimetry (DSC) analysis and was found to be 180.91 kJ/kg when melting and 181.22 kJ/kg when freezing.

Nanoencapsulation is still a novel technology. However, it would be interesting to see how the nanocapsules perform in comparison to microcapsules when integrated into building materials, with regards to both energy usage and structural stability in materials such as concrete. An issue that has not been discussed in the studies found is how nanocapsules will compare with respect to overall costs.

3.5.2.2 Adjustable Phase Change Temperature

Through most of the studies investigated in this review it is clear that the phase change temperature is one of the most important criteria for optimizing a PCM system. It is also clear that the optimal phase change temperature for summer conditions will not be optimal for winter conditions and vice versa. Some systems have been described where PCMs of different phase change temperatures are used to partly counter this issue. These systems work by adding different layers of PCMs with different phase change temperatures that can work optimally at different temperatures (Jin and Zhang, 2011; Pasupathy and Velraj, 2008). However, if this could be taken a step further, and a system with an adjustable phase change temperature could be created, the efficiency could be significantly improved. This could either be performed through a controllable or adaptive system and would enable a dynamic change of the phase change temperature in response to user preferences, different climates, and different seasons.

3.5.3 Further Reflections

3.5.3.1 Developing a Standard Test Scheme

With more and more commercial PCM products reaching the market, properties relevant for their use in construction have to be identified and tested. Mehling et al. (2006) showed that PCMs cannot be measured accurately enough by using standards that have been developed for other materials. These tests do not fully map all relevant effects around the phase change process due to PCMs’ low thermal conductivity and large melting enthalpy. This leads to an error in the stored heat as a function of temperature and may give errors that cannot be tolerated when evaluating the application of PCMs.

It has been pointed out that there is a lack of common test methods to compare and model the results of various PCMs (Agyenim et al., 2010; Shrestha et al., 2011). This makes it difficult to assess the suitability for each individual application. Accurate knowledge of a PCM’s enthalpy as a function of temperature or storage capacity at each temperature is important for design purposes (Lázaro et al., 2006).

Kousksou et al. (2014) reviewed the applications and challenges for energy storage systems, including PCMs. From this study it was found that there were still discrepancies between the literature and actual measurements of thermophysical properties of technical grade PCMs, further proving that a uniform test standard is needed.

It is important to note that the values measured for pure PCM compounds will not be representative when they are integrated into buildings. Encapsulation materials and combination with building materials will alter the response of the PCMs. It is therefore of importance that finished products aimed for the use in buildings are tested properly. Having a standard test scheme will also be beneficial for future work within field testing of PCMs as all values can be compared with respect to the same conditions.

There are still many issues regarding correct modeling of PCM behavior. Some of these issues include:

ent properly representing PCM subcooling

ent incorrect DSC data for whole building simulations

ent non-uniform PCMs cannot be tested in DSC

ent need for development of Δ-enthalpy charts for PCM-enhanced materials.

Subcooling, which is a common problem for PCMs based on salt hydrates, is still an effect that cannot be treated analytically or numerically. If salt hydrate PCMs become widespread, this is an effect that can have a major effect on a system’s performance, even at just 1 or 2°C of subcooling (Mehling et al., 2007).

The most common test methods used for evaluating the performance of PCMs to date are mentioned in the following sections. Further research into these methods may give a method that can be used as a common standard for testing PCMs in the future.

3.5.3.2 Differential Scanning Calorimetry

Currently, differential scanning calorimetry (DSC) is the most used method to determine a PCM’s thermophysical properties (Barreneche et al., 2013). DSC can only be used for testing small samples of PCMs. Through DSC it is possible to obtain melting temperatures and the heat of fusion. For this purpose, DSC is the most used method. However, for small samples, the degree of supercooling may be increased while the degree of phase segregation may be decreased, giving different results from actual use of the bulk materials used in practical systems (Zhang and Jiang, 1999). DSC can only be used on relatively uniform test specimens, which is not very realistic when testing building envelope products (Kosny et al., 2012a).

3.5.3.3 T-History

The T-history method allows for testing of the melting temperature, degree of supercooling, heat of fusion, specific heat, and thermal conductivity of several PCMs simultaneously (Zhang and Jiang, 1999).

Solé et al. (2013) wrote a review on the T-history method as it has been used to test PCMs up until recently. The aim of their study was to increase the consensus around the use of this method as it has several advantages when measuring PCMs. The T-history method is efficient at determining fusion enthalpy, specific heat, and thermal conductivity for large PCM samples. Nevertheless, there is still no commercial T-history equipment available yet.

3.5.3.4 Dynamic Heat Flow Apparatus

The dynamic heat flow apparatus method is based on the traditional heat flow meter apparatus method, which is used to determine steady-state heat transfer properties, thermal conductivity, and thermal resistance of flat slab specimens. The traditional heat flow apparatus method is already used in accordance with standards such as EN 12667, ISO 8301, and ASTM C518 to determine these properties for traditional insulation materials. When performing a dynamic heat flow test, the plates on each side of the specimen to be tested are held at the same temperature, and both plates are changed to a different temperature, whereas different temperatures are used in normal heat flow tests. By using the dynamic method, results were found to be more accurate when testing PCMs. Dynamic properties such as heat capacity profiles, peaks of melting and solidification cycles, and amount of subcooling were found to be relatively similar to those measured by DSC (Shukla et al., 2012).

3.5.3.5 Dynamic Hot Box

The dynamic hot box method can be used to simulate changes in temperature on the climate side of a test specimen. From this, the dynamic thermal characteristics of a building component can be found. When testing PCMs it is important to get as correct a picture as possible of their dynamic properties. With the dynamic hot box, the temperature is held constant until a steady-state is reached, then the temperature on one side is changed and results are measured until a new steady-state has been reached.

3.5.3.6 Dynamic Guarded Hot Plate

A study performed by Pomianowski et al. (2014) investigated the possibility of using a dynamic guarded hot plate apparatus to determine the specific heat capacity as a function of the temperature for PCM-incorporated concrete. The study describes the experimental setup and proposes various methods to calculate the specific heat capacity of PCM concrete. The advantage of this method is the possibility to attain a very small heating rate, which can imitate the temperature increase in real building constructions. That is, it can simulate realistic thermal conditions of the PCM in actual use. Though the study focuses on PCM concrete, it is pointed out that this experimental setup could also be used on various other PCM composites.

3.5.3.7 M-Value

PCMs are added to structures and lead to an increase in thermal mass. As PCMs have a specific temperature range, they have also been referred to as “smart thermal mass”. However, the phase change temperature of a PCM is not always absolute. It may happen over a small temperature interval and differ between melting and freezing (Fig. 3.16). This is a cause for confusion as there is no standard that specifies in which part of the phase change process this value should be stated. A new energy performance label for PCMs has been mentioned, i.e., the M-value, which should express the phase transition related enthalpy change of the PCM within a standardized range so that all values given for a PCM can be evaluated on an equal basis.

image
Figure 3.16 The current confusion around the correct enthalpy value for phase change (Kosny, 2011).

3.5.3.8 Environmental Impact Assessments

New materials that may be suitable for use as PCMs are continuously being investigated. As mentioned earlier, it is critical that PCMs for use in buildings are not hazardous or poisonous in any way. This also includes during production and handling after a PCM’s useful lifetime. It is important that possible new materials are evaluated with regard to waste from the production and necessary treatment when they are to be disposed. As there may be many different chemicals in use in PCMs, environmental assessments could be conducted to avoid unknown pollutants of any kind.

3.5.3.9 Expected Lifetime Predicament of Phase Change Materials

As a thawing/freezing cycle may happen thousands of times, some PCMs may be prone to deterioration over time. Gradual breakdown of the material will end up reducing the amount of latent heat storage per phase transition and may also affect the phase change temperature, and hence reduce the overall energy performance. Such deterioration is particularly common in inorganic PCMs, and to some extent in less pure organic PCMs. It is important that future possible PCMs are tested with regard to their aging over time. As of 2016 there is no standard method to do this. An aim for the future should be to develop a standard testing scheme that all commercial grade PCMs will have to follow when stating the lifetime of their products.

3.5.3.10 Quantifying the Effect of Phase Change Materials in Real-Life Buildings

Though the theory behind PCMs shows that they have the ability to reduce temperature fluctuations in buildings, there has been no study to prove an increase in overall thermal comfort or productivity. A way of measuring the effect and number of full cycles in different environments is needed.

To fully utilize the effect of PCMs, an overview of how they perform in different climates is needed. Ideally, a PCM should undergo a full phase change cycle once a day. However, the lack of guidelines for selecting PCMs for various climates is a source for uncertainty when used for building applications. It would be preferable if a detailed overview of the effect that could be expected for various PCMs in various climates was developed. It has been pointed out that it is difficult to select a PCM that functions optimally for every season and location.

From all the experiments and simulations reviewed in this study, one of the most important factors for implementing PCM enhanced materials is found to be selecting the correct phase change temperature. Other crucial aspects are the difficulty of selecting the correct PCM for specific climates and that the effect of PCMs will vary with varying climate conditions over the year.

The selection of a PCM based on a specific phase change temperature in one climate region will not be appropriate for another. Important factors that govern the selection of a PCM include phase change temperature, local climate, type of PCM and design and orientation of the building (Pasupathy and Velraj, 2008). Several other authors have also highlighted the need for guidelines to select PCMs for specific and different climates (Kosny et al., 2013; Ascione et al., 2014).

From the various manufacturers that have been investigated in this study a large amount of reference projects have been found. However, to the authors’ knowledge, there have been no studies to evaluate and quantify the effect of PCMs in these projects. Although every project has been deemed to increase energy efficiency and thermal comfort, monitoring projects from real life buildings should be conducted and evaluated so the beneficial effects of PCMs can be documented and shared.

3.5.3.11 Investigating Payback Times for Various Systems

As PCMs offer a decrease in overall energy usage, in many cases through minimal interference with the original structure, it is highly relevant to know the expected payback time of the initial investment. To the authors’ knowledge, little research has been carried out on this topic. This may be due to several factors, the most important one being that there is still a lack of knowledge on how PCMs actually perform in real life constructions in all sorts of various climates.

There is still much uncertainty regarding the overall effect of PCMs. Showing the long-term economical benefits of selecting PCMs may help to increase interest among customers to use PCMs. Should the results of such analyses show unfavorable payback times, it may also help to drive the research and development forward towards more cost-efficient solutions.

3.5.3.12 Development of Advanced Building Envelopes

PCMs for building applications may be viewed as part of the development of new materials and solutions for the building envelope in energy- efficient buildings. A seamless and robust integration of these materials and technologies with the PCM parts is important in order to ensure energy-efficient and durable solutions during the whole anticipated lifetime. Although not thermal insulation materials in themselves, PCMs have already been mentioned as part of the thermal building envelope (Kalnæs and Jelle, 2015). State-of-the-art thermal insulation material solutions include vacuum insulation panels (VIP) (Baetens et al., 2010a; Brunner et al., 2014; Grynning et al., 2011; Haavi et al., 2012; Jelle et al., 2013a; Johansson et al., 2014; Kalnæs and Jelle, 2014; Sveipe et al., 2011; Tenpierik, 2009; Wegger et al., 2011), gas-filled panels (GFP) (Baetens et al., 2010c), and aerogels (Baetens et al., 2011; Jelle et al., 2015c; Koebel et al., 2012; Levy and Zayat, 2015). Furthermore, investigations with superinsulation materials (SIM), e.g., nanoinsulation materials (NIM), are ongoing (Baetens et al., 2010a; Jelle, 2011a; Jelle et al., 2009, 2010, 2011a), including experimental attempts with hollow silica nanospheres (HSNS) (Gao et al., 2012, 2013, 2014a, 2015; Grandcolas et al., 2013; Jelle et al., 2011a, 2013b, 2014a, 2015a; Sandberg et al., 2013). Life-cycle assessments of these HSNS have also been performed (Gao et al., 2013, 2014a; Schlanbusch et al., 2014a,b). Moreover, experimental investigations with aerogel–concrete mixtures have also been conducted (Gao et al., 2014b; Ng et al., 2015a,b, 2016).

In connection to windows and fenestration in general, there is a lot of development in several fields (Jelle et al., 2012a), including, e.g., smart and electrochromic windows (Granqvist, 1995, 2005, 2012; Jelle, 1993, 2013a, 2015; Jelle and Hagen, 1993, 1998, 1999a,b; Jelle et al., 1992a,b, 1993a,b,c, 1998, 2007; Lampert, 1998, 2004; Mortimer et al., 2015; Ribeiro and Mortimer, 2015), window pane spacer materials (Van Den Bergh et al., 2013; Asphaug et al., 2016), low-emissivity coatings (Jelle et al., 2015b), window frame technologies (Gustavsen et al., 2008, 2011), self-cleaning surfaces (Midtdal and Jelle, 2013; Jelle, 2013b), solar cell glazing products (Jelle, 2016; Jelle and Breivik, 2012a,b), various fenestration aspects (Jelle et al., 2012a), energy optimization (Grynning et al., 2013b), translucent aerogel window studies (Gao et al., 2014c, 2016; Ihara et al., 2015a,b,c), and new lightweight aerogel window pane glass materials (Gao et al., 2014d).

Building integrated photovoltaics (BIPV) will probably become more and more important in the years to come, thus providing energy-efficient buildings with an electricity power generating source (Breivik et al., 2013; Jelle, 2016; Jelle and Breivik, 2012a,b; Jelle et al., 2012b), where nanotechnology plays a crucial role in the development of new solar cell materials (Jelle, 2016; Jelle et al., 2012b), also including the challenge of removing or avoiding snow and ice formation on the solar cell surfaces, e.g., by the formation and application of superhydrophobic water-repelling and icephobic anti-icing surfaces (Dash et al., 2012; Eberle et al., 2014; Jelle, 2013b, 2016; Kreder et al., 2016; Schutzius et al., 2015; Xiao and Chaudhuri, 2012; Zhang and Lv, 2015).

To ensure energy-efficient and durable buildings all kinds of different aspects have to be considered, where satisfactory resistance towards weather exposure is one important aspect, e.g., from shielding of fresh air ventilation intakes (Jelle and Noreng, 2012) to wind-driven raintightness of BIPV (Breivik et al., 2013; Fasana and Nelva, 2013). Contrary to the heavy climate exposure that walls and roofs experience, the climate exposure for the ground may seem much less harsh; nevertheless, the ground materials and solutions must prevent, e.g., penetration of air (leakage), moisture, and radon (Jelle et al., 2011b; Jelle, 2012a; Pacheco-Torgal, 2012), thus also comprising important health issues (fungi growth and radioactive radiation exposure due to moisture damages and radon ingress, respectively). Air, moisture, and radon barriers towards the ground may be combined into one barrier depending on the placement location of the barrier, and may also incorporate a low-emissivity surface coating (Jelle et al., 2015b) to decrease the heat loss towards the ground.

Miscellaneous advantages and drawbacks of applying nanotechnology in the construction sector are reviewed by Pacheco-Torgal and Jalali (2011). Certainly, as one in the coming years will experience many new and innovative materials and solutions based on nanotechnology, one also has to address any toxicity issues to avoid any health and safety hazards. That also goes for the development of new PCMs and other technologies to be used in the building envelope, e.g., the ones mentioned in the above. Several of these new materials will probably be developed by research groups within the European Commission Horizon 2020 program, where the utilization of nanotechnology also will play an important role (Pacheco-Torgal, 2014).

In general, it is of major importance to investigate the durability of building materials and components, also newly developed ones, e.g., by carrying out accelerated climate aging in the laboratory (Jelle, 2011b, 2012b, 2013c; Jelle and Nilsen, 2011; Jelle et al., 2012c). Thus, performing a robustness assessment of these materials and components may also be found to be beneficial (Jelle et al., 2014b). That is, a durability and robustness evaluation of the new PCMs should be carried out. Applying all these different technologies with their specific properties fulfilling various functions, also including, e.g., PCMs, one may start denoting these building envelopes as multifunctional building envelopes.

Several of the materials to be utilized in advanced and multifunctional building envelopes may be denoted as dynamic materials and further categorized as either adaptive or controllable materials. Photochromic and thermochromic materials that change their solar radiation transmittance according to the ambient solar radiation exposure and temperature, respectively, are examples of adaptive materials. Electrochromic materials that change their solar radiation transmittance according to an applied voltage represent an example of controllable materials, i.e., properties controlled by human beings. PCMs are categorized as adaptive materials. However, if the phase change temperature could be made adjustable either by controllable or adaptive means, especially by the former, one may start talking about semicontrollable PCMs, thus extending the usefulness through increased flexibility and increasing the energy efficiency of the PCM system.

3.6 Conclusions

Incorporating phase change materials (PCM) into a building enables a more dynamic use of energy. Due to the storage capabilities of PCMs, excess heat can be stored during warm periods and released during cold periods. It may also work the other way around, storing cold energy and using it for free cooling systems in warm periods. The benefits of using PCMs in buildings mainly revolve around a decrease in energy usage along with a peak load shifting of energy required for heating or cooling and an increase in thermal comfort by decreasing temperature fluctuations.

Commercial products have been developed and released on the market with some success. What makes PCMs particularly interesting is the fact that many PCM solutions, e.g., wallboards and floor tiles, can be added to the construction with fairly little alteration to the current way of building. However, there is a wide range of materials that can be used as PCMs, and identifying the correct PCM for the specific application and for the specific climate conditions is an area that needs further research. Calculations of payback periods for PCM installations are also needed to further validate the use of the PCM technology.

Though solutions have been tested to increase thermal conductivity for more effective absorbance and discharge cycles, this has come at the cost of a lower latent heat storage per unit weight and unit volume, hence giving the PCMs less storage potential. Fire safety is still an issue for organic PCMs, though here as well, solutions that show promise have been introduced.

The PCM technology seems promising; however, there are still some hurdles that need to be overcome for a large-scale application of this technology. Standards that state test methods and can help identify the correct PCMs for various climates, to enable proper cycling and optimization of PCM systems, are needed. Research into new PCM technologies is also of major importance, e.g., the possibility of having a dynamically adjustable and even controllable phase change temperature.

Acknowledgments

This work has been supported by the Research Council of Norway and several partners through The Research Centre on Zero Emission Buildings (ZEB).

References

1. Agyenim F, Hewitt N, Eames P, Smyth M. A review of materials, heat transfer and phase change problem formulation for latent heat thermal energy storage systems (LHTESS). Renew Sustain Energy Rev. 2010;14:615–628.

2. Alawadhi EM. Using phase change materials in window shutter to reduce the solar heat gain. Energy Build. 2012;47:421–429.

3. Alkan C. Enthalpy of melting and solidification of sulfonated paraffins as phase change materials for thermal energy storage. Thermochim Acta. 2006;451:126–130.

4. Alqallaf HJ, Alawadhi EM. Concrete roof with cylindrical holes containing PCM to reduce the heat gain. Energy Build. 2013;61:73–80.

5. Al-Saadi SN, Zhai Z. Modeling phase change materials embedded in building enclosure: a review. Renew Sustain Energy Rev. 2013;21:659–673.

6. Ansuini R, Larghetti R, Giretti A, Lemma M. Radiant floors integrated with PCM for indoor temperature control. Energy Build. 2011;43:3019–3026.

7. Ascione F, Bianco N, De Masi RF, de' Rossi F, Vanoli GP. Energy refurbishment of existing buildings through the use of phase change materials: energy savings and indoor comfort in the cooling season. Appl Energy. 2014;113:990–1007.

8. Asphaug SK, Jelle BP, Gullbrekken L, Uvsløkk S. Accelerated ageing and durability of double-glazed sealed insulating window panes and impact on heating demand in buildings. Energy Build. 2016;116:395–402.

9. Babaei H, Keblinski P, Khodadadi JM. Improvement in thermal conductivity of paraffin by adding high aspect-ratio carbon-based nano-fillers. Phys Lett A. 2013;377:1358–1361.

10. Baetens R, Jelle BP, Thue JV, et al. Vacuum insulation panels for building applications: a review and beyond. Energy Build. 2010a;42:147–172.

11. Baetens R, Jelle BP, Gustavsen A. Phase change materials for building applications: a state-of-the-art review. Energy Build. 2010b;42:1361–1368.

12. Baetens R, Jelle BP, Gustavsen A, Grynning S. Gas-filled panels for building applications: a state-of-the-art review. Energy Build. 2010c;42:1969–1975.

13. Baetens R, Jelle BP, Gustavsen A. Properties, requirements and possibilities of smart windows for dynamic daylight and solar energy control in buildings: a state-of-the-art review. Sol Energy Mater Sol Cells. 2010d;94:87–105.

14. Baetens R, Jelle BP, Gustavsen A. Aerogel insulation for building applications: a state-of-the-art review. Energy Build. 2011;43:761–769.

15. Barreneche C, Solé A, Miró L, Martorell I, Fernández AI, Cabeza LF. Study on differential scanning calorimetry with two operation modes and organic and inorganic phase change material (PCM). Thermochim Acta. 2013;553:23–26.

16. BASF, 2013. BASF Micronal. Available from: <http://www.micronal.de/portal/basf/ien/dt.jsp?setCursor=1_290798> (accessed 01.12.2013).

17. Becker R. Improving thermal and energy performance of buildings in summer with internal phase change materials. J Build Phys. 2014;37:296–324.

18. Bourne S, Novoselac A. PCM-based high-density thermal storage systems for residential and small commercial retrofit applications. Procedia Eng. 2015;121:536–543.

19. Breivik C, Jelle BP, Time B, et al. Large-scale experimental wind-driven rain exposure investigations of building integrated photovoltaics. Sol Energy. 2013;90:179–187.

20. Brunner S, Wakili KG, Stahl T, Binder B. Vacuum insulation panels for building applications—continuous challenges and developments. Energy Build. 2014;85:592–596.

21. Cabeza LF, Castellón C, Nogués M, Medrano M, Leppers R, Zubillaga O. Use of microencapsulated PCM in concrete walls for energy savings. Energy Build. 2007;39:113–119.

22. Cabeza LF, Castell A, Barreneche C, de Gracia A, Fernández AI. Materials used as PCM in thermal energy storage in buildings: a review. Renew Sustain Energy Rev. 2011;15:1675–1695.

23. Cao L, Tang F, Fang G. Synthesis and characterization of microencapsulated paraffin with titanium dioxide shell as shape-stabilized thermal energy storage materials in buildings. Energy Build. 2015;72:31–37.

24. Cao, S., Gustavsen, A., Uvsløkk, S., Jelle, B.P., Gilbert, J., Maunuksela, J., 2010. The effect of wall-integrated phase change material panels on the indoor air and wall temperature—hot box experiments. In: Proceedings of the Renewable Energy Research Conference—Renewable Energy Beyond 2020, Trondheim, Norway, June 7–8.

25. Chen ZH, Yu F, Zeng XR, Zhang ZG. Preparation, characterization and thermal properties of nanocapsules containing phase change material n-dodecanol by miniemulsion polymerization with polymerizable emulsifier. Appl Energy. 2012;91:7–12.

26. Cho JS, Kwon A, Cho CG. Microencapsulation of octadecane as a phase-change material by interfacial polymerization in an emulsion system. Colloid Polym Sci. 2002;280:260–266.

27. Chwieduk DA. Dynamics of external wall structures with a PCM (phase change materials) in high latitude countries. Energy. 2013;59:301–313.

28. Cosella-Dörken, 2014a.Delta-Cool 21 used in a floor application of the north house at the US department of energy solar decathlon, October 2009. Available from: <http://www.cosella-dorken.com/bvf-ca-en/projects/pcm/northhouse.php> (accessed 08.01.2014).

29. Cosella-Dörken, 2014b. Delta-Cool 28 used in a glass facade application on a zero energy office building in Kempen, CH. Available from: <http://www.cosella-dorken.com/bvf-ca-en/projects/pcm/kempen.php> (accessed 08.01.2014).

30. Cui Y, Xie J, Liu J, Pan S. Review of phase change materials integrated in building walls for energy saving. Procedia Eng. 2015;121:763–770.

31. Dash S, Alt MT, Garimella SV. Hybrid surface design for robust superhydrophobicity. Langmuir. 2012;28:9606–9615.

32. Diaconu BM. Thermal energy savings in buildings with PCM-enhanced envelope: influence of occupancy pattern and ventilation. Energy Build. 2011;43:101–107.

33. Dieckmann JH. Latent Heat Storage in Concrete Kaiserslautern, Germany: Technische Universität Kaiserslautern; 2013; <http://www.eurosolar.org/new/pdfs_neu/Thermal/IRES2006_Dieckmann.pdf> (accessed 31.10.2013).

34. Eberle P, Tiwari MK, Maitra T, Poulikakos D. Rational nanostructuring of surfaces for extraordinary icephobicity. Nanoscale. 2014;6:4874–4881.

35. European Union. Directive 2012/27/EU of the European parliament and of the council of 25 October 2012 on the energy efficiency. Off J Eur Union. 2012;Vol. 55:1–57 L 315.

36. Evola G, Marletta L, Sicurella F. A methodology for investigating the effectiveness of PCM wallboards for summer thermal comfort in buildings. Build Environ. 2013;59:517–527.

37. Fan L, Khodadadi JM. Thermal conductivity enhancement of phase change materials for thermal energy storage: a review. Renew Sustain Energy Rev. 2011;15:24–46.

38. Farid MM, Khudhair AM, Razack SAK, Al-Hallaj S. A review on phase change energy storage: materials and applications. Energy Convers Manag. 2004;45:1597–1615.

39. Fasana S, Nelva R. Improvement of the water resistancy in the integration of photovoltaic panels on traditional roofs. Constr Build Mater. 2013;48:1081–1091.

40. Fiorentini M, Cooper P, Ma Z. Development and optimization of an innovative HVAC system with integrated PVT and PCM thermal storage for a net-zero energy retrofitted house. Energy Build. 2015;94:21–32.

41. Gao T, Sandberg LIC, Jelle BP, Gustavsen A. Nano insulation materials for energy efficient buildings: a case study on hollow silica nanospheres. In: Mendez-Vilas A, ed. Fuelling the Future: Advances in Science and Technologies for Energy Generation, Transmission and Storage. BrownWalker Press 2012;535–539.

42. Gao T, Jelle BP, Sandberg LIC, Gustavsen A. Monodisperse hollow silica nanospheres for nano insulation materials: synthesis, characterization, and life cycle assessment. ACS Appl Mater Interface. 2013;5:761–767.

43. Gao T, Sandberg LIC, Jelle BP. Nano insulation materials: synthesis and life cycle assessment. Procedia CIRP. 2014a;15:490–495.

44. Gao T, Jelle BP, Gustavsen A, Jacobsen S. Aerogel-incorporated concrete: an experimental study. Constr Build Mater. 2014b;52:130–136.

45. Gao T, Jelle BP, Ihara T, Gustavsen A. Insulating glazing units with silica aerogel granules: the impact of particle size. Appl Energy. 2014c;128:27–34.

46. Gao T, Jelle BP, Gustavsen A, He J. Lightweight and thermally insulating aerogel glass materials. Appl Phys A: Mater Sci Process. 2014d;117:799–808.

47. Gao T, Jelle BP, Sandberg LIC, Gustavsen A. Thermal conductivity of monodisperse silica nanospheres. J Porous Media. 2015;18:941–947.

48. Gao T, Ihara T, Grynning S, Jelle BP, Lien AG. Perspective of aerogel glazings in energy efficient buildings. Build Environ. 2016;95:405–413.

49. Giro-Paloma J, Martínez M, Cabeza LF, Fernández AI. Types, methods, techniques, and applications for microencapsulated phase change materials (MPCM): a review. Renew Sustain Energy Rev. 2016;53:1059–1075.

50. GlassX, 2015. GlassX crystal–Das Glas, das speichert, wärmt und kühlt. Available from: <http://glassx.ch/index.php?id=564> (Broschuere_klein_online.pdf) (accessed 22.01.2015).

51. Goia F, Perino M, Serra V. Improving thermal comfort conditions by means of PCM glazing systems. Energy Build. 2013;60:442–452.

52. Goia F, Perino M, Serra V. Experimental analysis of the energy performance of a full-scale PCM glazing prototype. Sol Energy. 2014;100:217–233.

53. Grandcolas, M., Etienne, G., Tilset, B.G., Gao, T., Sandberg, L.I.C., Gustavsen, A., et al., 2013. Hollow silica nanospheres as a superinsulating material. In: Proceedings of the 11th International Vacuum Insulation Symposium (IVIS 2013), Dübendorf, Zürich, Switzerland, September 19–20, pp. 43–44.

54. Granqvist CG. Handbook of Inorganic Electrochromic Materials Amsterdam: Elsevier; 1995.

55. Granqvist CG. Electrochromic devices. J Eur Ceram Soc. 2005;25:2907–2912.

56. Granqvist CG. Oxide electrochromics: an introduction to devices and materials. Sol Energy Mater Sol Cells. 2012;99:1–13.

57. Grynning S, Jelle BP, Uvsløkk S, et al. Hot box investigations and theoretical assessments of miscellaneous vacuum insulation panel configurations in building envelopes. J Build Phys. 2011;34:297–324.

58. Grynning S, Goia F, Rognvik E, Time B. Possibilities for characterization of a PCM window system using large scale measurements. Int J Sustain Built Environ. 2013a;2:56–64.

59. Grynning S, Gustavsen A, Time B, Jelle BP. Windows in the buildings of tomorrow: energy losers or energy gainers? Energy Build. 2013b;61:185–192.

60. Gustavsen A, Arasteh D, Jelle BP, Curcija C, Kohler C. Developing low-conductance window frames: capabilities and limitations of current window heat-transfer design tools. J Build Phys. 2008;32:131–153.

61. Gustavsen A, Grynning S, Arasteh D, Jelle BP, Goudey H. Key elements of and materials performance targets for highly insulating window frames. Energy Build. 2011;43:2583–2594.

62. Haavi T, Jelle BP, Gustavsen A. Vacuum insulation panels in wood frame wall constructions with different stud profiles. J Build Phys. 2012;36:212–226.

63. Halford CK, Boehm RF. Modeling of phase change material peak load shifting. Energy Build. 2007;39:298–305.

64. Hawlader MNA, Uddin MS, Khin MM. Microencapsulated PCM thermal-energy storage system. Appl Energy. 2003;74:195–202.

65. Ihara T, Grynning S, Gao T, Gustavsen A, Jelle BP. Impact of convection on thermal performance of aerogel granulate glazing systems. Energy Build. 2015a;88:165–173.

66. Ihara T, Gao T, Grynning S, Jelle BP, Gustavsen A. Aerogel granulate glazing facades and their application potential from an energy saving perspective. Appl Energy. 2015b;142:179–191.

67. Ihara T, Jelle BP, Gao T, Gustavsen A. Aerogel granule aging driven by moisture and solar radiation. Energy Build. 2015c;103:238–248.

68. Ismail KAR, Salinas CT, Henriquez JR. Comparison between PCM filled glass windows and absorbing gas filled windows. Energy Build. 2008;40:710–719.

69. Jain L, Sharma SD. Phase change materials for day lighting and glazed insulation in buildings. J Eng Sci Technol. 2009;4:322–327.

70. Jelle, B.P., 1993. Electrochemical and spectroscopic studies of electrochromic materials. Ph.D. thesis, Department of Applied Electrochemistry, The Norwegian Institute of Technology, Trondheim, Norway.

71. Jelle BP. Traditional, state-of-the-art and future thermal building insulation materials and solutions—properties, requirements and possibilities. Energy Build. 2011a;43:2549–2563.

72. Jelle, B.P., 2011b. Evaluation of building products by conducting accelerated climate ageing in the laboratory. In: Proceedings of the 12th International Conference on Durability of Building Materials and Components, Porto, Portugal, April, 12–15, pp. 311–319.

73. Jelle BP. Development of a model for radon concentration in indoor air. Sci Total Environ. 2012a;416:343–350.

74. Jelle BP. Accelerated climate ageing of building materials, components and structures in the laboratory. J Mater Sci. 2012b;47:6475–6496.

75. Jelle BP. Solar radiation glazing factors for window panes, glass structures and electrochromic windows in buildings—measurement and calculation. Sol Energy Mater Sol Cells. 2013a;116:291–323.

76. Jelle BP. The challenge of removing snow downfall on photovoltaic solar cell roofs in order to maximize solar energy efficiency—research opportunities for the future. Energy Build. 2013b;67:334–351.

77. Jelle, B.P., 2013c. The role of accelerated climate ageing of building materials, components and structures in the laboratory. In: Proceedings of the 7th Nordic Conference on Construction Economics and Organisation 2013, Trondheim, Norway, June 12–14, pp. 111–122.

78. Jelle BP. Electrochromic smart windows for dynamic daylight and solar energy control in buildings. In: Mortimer RJ, Rosseinsky DR, Monk PMS, eds. Electrochromic Materials and Devices. Wiley-VCH 2015;419–502.

79. Jelle BP. Building integrated photovoltaics: a concise description of the current state of the art and possible research pathways. Energy. 2016;9:1–30 Article no. 21.

80. Jelle BP, Hagen G. Transmission spectra of an electrochromic window based on polyaniline, prussian blue and tungsten oxide. J Electrochem Soc. 1993;140:3560–3564.

81. Jelle BP, Hagen G. Electrochemical multilayer deposition of polyaniline and prussian blue and their application in solid state electrochromic windows. J Appl Electrochem. 1998;28:1061–1065.

82. Jelle BP, Hagen G. Performance of an electrochromic window based on polyaniline, prussian blue and tungsten oxide. Sol Energy Mater Sol Cells. 1999a;58:277–286.

83. Jelle BP, Hagen G. Correlation between light absorption and electric charge in solid state electrochromic windows. J Appl Electrochem. 1999b;29:1103–1110.

84. Jelle BP, Nilsen T-N. Comparison of accelerated climate ageing methods of polymer building materials by attenuated total reflectance Fourier transform infrared radiation spectroscopy. Constr Build Mater. 2011;25:2122–2132.

85. Jelle BP, Breivik C. State-of-the-art building integrated photovoltaics. Energy Procedia. 2012a;20:68–77.

86. Jelle BP, Breivik C. The path to the building integrated photovoltaics of tomorrow. Energy Procedia. 2012b;20:78–87.

87. Jelle BP, Noreng K. Shielding fresh air ventilation intakes. ASHRAE J. 2012; April, 34–42.

88. Jelle BP, Hagen G, Hesjevik SM, Ødegård R. Transmission through an electrochromic window based on polyaniline, tungsten oxide and a solid polymer electrolyte. Mater Sci Eng B. 1992a;B13:239–241.

89. Jelle BP, Hagen G, Ødegård R. Transmission spectra of an electrochromic window based on polyaniline, tungsten oxide and a solid polymer electrolyte. Electrochim Acta. 1992b;37:1377–1380.

90. Jelle BP, Hagen G, Sunde S, Ødegård R. Dynamic light modulation in an electrochromic window consisting of polyaniline, tungsten oxide and a solid polymer electrolyte. Synth Met. 1993a;54:315–320.

91. Jelle BP, Hagen G, Hesjevik SM, Ødegård R. Reduction factor for polyaniline films on ITO from cyclic voltammetry and visible absorption spectra. Electrochim Acta. 1993b;38:1643–1647.

92. Jelle BP, Hagen G, Nødland S. Transmission spectra of an electrochromic window consisting of polyaniline, prussian blue and tungsten oxide. Electrochim Acta. 1993c;38:1497–1500.

93. Jelle BP, Hagen G, Birketveit Ø. Transmission properties for individual electrochromic layers in solid state devices based on polyaniline, prussian blue and tungsten oxide. J Appl Electrochem. 1998;28:483–489.

94. Jelle BP, Gustavsen A, Nilsen T-N, Jacobsen T. Solar material protection factor (SMPF) and solar skin protection factor (SSPF) for window panes and other glass structures in buildings. Sol Energy Mater Sol Cells. 2007;91:342–354.

95. Jelle, B.P., Gustavsen, A., Baetens, R., 2009. Beyond vacuum insulation panels—How may it be achieved? In: Proceedings of the 9th International Vacuum Insulation Symposium (IVIS 2009), London, England, September, 17–18.

96. Jelle BP, Gustavsen A, Baetens R. The path to the high performance thermal building insulation materials and solutions of tomorrow. J Build Phys. 2010;34:99–123.

97. Jelle, B.P., Tilset, B.G., Jahren, S., Gao, T., Gustavsen, A., 2011a. Vacuum and nanotechnologies for the thermal insulation materials of beyond tomorrow—from concept to experimental investigations. In: Proceedings of the 10th International Vacuum Insulation Symposium (IVIS 2011), Ottawa, Canada, September, 15–16, pp. 171–178.

98. Jelle BP, Noreng K, Erichsen TH, Strand T. Implementation of radon barriers, model development and calculation of radon concentration in indoor air. J Build Phys. 2011b;34:195–222.

99. Jelle BP, Hynd A, Gustavsen A, Arasteh D, Goudey H, Hart R. Fenestration of today and tomorrow: a state-of-the-art review and future research opportunities. Sol Energy Mater Sol Cells. 2012a;96:1–28.

100. Jelle BP, Breivik C, Røkenes HD. Building integrated photovoltaic products: a state-of-the-art review and future research opportunities. Sol Energy Mater Sol Cells. 2012b;100:69–96.

101. Jelle BP, Nilsen T-N, Hovde PJ, Gustavsen A. Accelerated climate aging of building materials and their characterization by Fourier transform infrared radiation analysis. J Build Phys. 2012c;36:99–112.

102. Jelle, B.P., Sveipe, E., Wegger, E., Uvsløkk, S., Grynning, S., Thue, J.V., et al., 2013a. Moisture robustness during retrofitting of timber frame walls with vacuum insulation panels: experimental and theoretical studies. In: de Freitas, V.P., Delgado, J.P.Q. et al. (Eds.), Hygrothermal Behavior, Building Pathology and Durability, Book Series “Building Pathology and Rehabilitation”. Springer, Vol. 1, pp. 183–210.

103. Jelle, B.P., Gao, T., Tilset, B.G., Sandberg, L.I.C., Grandcolas, M., Simon, C., et al., 2013b. Experimental pathways for achieving superinsulation through nano insulation materials. In: Proceedings of the 11th International Vacuum Insulation Symposium (IVIS 2013), Dübendorf, Zürich, Switzerland, September, 19–20, pp. 99–100.

104. Jelle BP, Gao T, Sandberg LIC, Tilset BG, Grandcolas M, Gustavsen A. Thermal superinsulation for building applications—from concepts to experimental investigations. Int J Struct Anal Des. 2014a;1:43–50.

105. Jelle BP, Sveipe E, Wegger E, et al. Robustness classification of materials, assemblies and buildings. J Build Phys. 2014b;37:213–245.

106. Jelle, B.P., Gao, T., Sandberg, L.I.C., Ng, S., Tilset, B.G., Grandcolas, M., et al., 2015a. Development of nano insulation materials for building constructions. In: Proceedings of 5th International Symposium on Nanotechnology in Construction (NICOM5), Chicago, IL, May, 24–26, pp. 429–434.

107. Jelle BP, Kalnæs SE, Gao T. Low-emissivity materials for building applications: a state-of-the-art review and future research perspectives. Energy Build. 2015b;96:329–356.

108. Jelle BP, Baetens R, Gustavsen A. Aerogel insulation for building applications. In: Wiley-VCH 2015c;1385–1412. Levy D, Zayat M, eds. The Sol–Gel Handbook. Vol. 3.

109. Ji P, Sun H, Zhong Y, Feng W. Improvement of the thermal conductivity of a phase change material by the functionalized carbon nanotubes. Chem Eng Sci. 2012;81:140–145.

110. Jiang J, Zhu Y, Ma A, et al. Preparation and performance of bulk porous Al foams impregnated with phase-change-materials for thermal storage. Prog Nat Sci Mater Int. 2012;22:440–444.

111. Jin X, Zhang X. Thermal analysis of a double layer phase change material floor. Appl Therm Eng. 2011;31:1576–1581.

112. Johansson P, Geving S, Hagentoft C-E, et al. Interior insulation retrofit of a historical brick wall using vacuum insulation panels: hygrothermal numerical simulations and laboratory investigations. Build Environ. 2014;79:31–45.

113. Jurkowska M, Szczygieł I. Review on properties of microencapsulated phase change materials slurries (mPCMS). Appl Therm Eng. 2016;98:365–373.

114. Kalnæs SE, Jelle BP. Vacuum insulation panel products: a state-of-the-art review and future research pathways. Appl Energy. 2014;116:355–375.

115. Kalnæs SE, Jelle BP. Phase change materials and products for building applications: a state-of-the-art review and future research opportunities. Energy Build. 2015;94:150–176.

116. Karaipekli A, Sari A. Capric-myristic acid/expanded perlite composite as form-stable phase change material for latent heat thermal energy storage. Renew Energy. 2008;33:2599–2605.

117. Kaygusuz K, Alkan C, Sari A, Uzun O. Encapsulated fatty acids in an acrylic resin as shape-stabilized phase change materials for latent heat thermal energy storage. Energy Source. 2008;30:1050–1059.

118. Kenisarin M, Mahkamov K. Solar energy storage using phase change materials. Renew Sustain Energy Rev. 2007;11:1913–1965.

119. Kenisarin M, Mahkamov K. Passive thermal control in residential buildings using phase change materials. Renew Sustain Energy Rev. 2016;55:371–398.

120. Khadiran T, Hussein MZ, Zainal Z, Rusli R. Encapsulation techniques for organic phase change materials as thermal energy storage medium: a review. Sol Energy Mater Sol Cells. 2015;143:78–98.

121. Khudhair AM, Farid MM. A review on energy conservation in building applications with thermal storage by latent heat using phase change materials. Energy Convers Manag. 2004;45:263–275.

122. Kibria MA, Anisur MR, Mahfuz MH, Saidur R, Metselaar IHSC. A review on thermophysical properties of nanoparticle dispersed phase change materials. Energy Convers Manag. 2015;95:69–89.

123. Koebel M, Rigacci A, Achard P. Aerogel-based thermal superinsulation: an overview. J Sol–Gel Sci Technol. 2012;63:315–339.

124. Konuklu Y, Ostry M, Paksoy HO, Charvat P. Review on using microencapsulated phase change materials (PCM) in building applications. Energy Build. 2015;106:134–155.

125. Kosny, J., 2011. Opportunities to apply phase change materials to building enclosures. In: Building America, Cambridge, MA, USA, November 11. <http://apps1.eere.energy.gov/buildings/publications/pdfs/building_america/webinar_pcm_enclosures_20111111.pdf> (accessed 24.02.2016).

126. Kosny J, Kossecka E, Brzezinski A, Tleoubaev A, Yarbrough D. Dynamic thermal performance analysis of fiber insulations containing bio-based phase change materials (PCMs). Energy Build. 2012a;52:122–131.

127. Kosny J, Biswas K, Miller W, Kriner S. Field thermal performance of naturally ventilated solar roof with PCM heat sink. Sol Energy. 2012b;86:2504–2514.

128. Kosny, J., Shukla, N., Fallahi, A., 2013. Cost analysis of sample phase change material-enhanced building envelopes in southern U.S. climates. In: Building Technologies Program. U.S. Department of Energy, DOE/GO-102013-103692, January.

129. Kousksou T, Bruel P, Jamil A, Rhafiki TE, Zeraouli Y. Energy storage: applications and challenges. Sol Energy Mater Sol Cells. 2014;120:59–80.

130. Kreder MJ, Alvarenga J, Kim P, Aizenberg J. Design of anti-icing surfaces: smooth, textured or slippery? Nat Rev Mater. 2016;1:1–15 Article no. 15003.

131. Kuznik F, Virgone J. Experimental assessment of a phase change material for wall building use. Appl Energy. 2009;86:2038–2046.

132. Kuznik F, Virgone J, Roux JJ. Energetic efficiency of room wall containing PCM wallboard: a full-scale experimental investigation. Energy Build. 2008;40:148–156.

133. Kuznik F, David D, Johannes K, Roux JJ. A review on phase change materials integrated in building walls. Renew Sustain Energy Rev. 2011a;15:379–391.

134. Kuznik F, Virgone J, Johannes K. In-situ study of thermal comfort enhancement in a renovated building equipped with phase change material wallboard. Renew Energy. 2011b;36:1458–1462.

135. Lampert CM. Smart switchable glazing for solar energy and daylight control. Sol Energy Mater Sol Cells. 1998;52:207–221.

136. Lampert CM. Chromogenic smart materials. Mater Today. 2004;7:28–35.

137. Lan L, Wargocki P, Lian Z. Quantitative measurements of productivity loss due to thermal discomfort. Energy Build. 2011;43:1057–1062.

138. Lane GA. Phase change thermal storage materials. In: Guyer C, ed. Hand Book of Thermal Design. McGraw Hill Book Co. 1989.

139. Latibari ST, Mehrali M, Mehrali M, Mahlia TMI, Metselaar HSC. Synthesis, characterization and thermal properties of nanoencapsulated phase change materials via sol–gel method. Energy. 2013;61:664–672.

140. Lázaro A, Günther E, Mehling H, Hiebler S, Marín JM, Zalba B. Verification of a T-history installation to measure enthalpy versus temperature curves of phase change materials. Meas Sci Technol. 2006;17:2168–2174.

141. Levy D, Zayat M, eds. The Sol–Gel Handbook. Wiley-VCH 2015.

142. Liang C, Lingling X, Hongbo S, Zhibin Z. Microencapsulation of butyl stearate as a phase change material by interfacial polycondensation in a polyurea system. Energy Convers Manag. 2009;50:723–729.

143. Ling TC, Poon CS. Use of phase change materials for thermal energy storage in concrete: an overview. Constr Build Mater. 2013;46:55–62.

144. Liu C, Rao Z, Zhao J, Huo Y, Li Y. Review on nanoencapsulated phase change materials: preparation, characterization and heat transfer enhancement. Nano Energy. 2015;13:814–826.

145. Ma Z, Lin W, Sohel MI. Nano-enhanced phase change materials for improved building performance. Renew Sustain Energy Rev. 2016;58:1256–1268.

146. Markets and Markets, 2013. Advanced phase change material (PCM) market, by type (inorganic, organic and bio-based) & application (building & construction, cold storage, HVAC, textile, thermal energy storage, electronics)—global trends & forecast to 2018.

147. Mehling, H., Ebert, H.-P., Schossig, P., 2006. Development of standards for materials testing and quality control of PCM. In: Proceedings of the 7th IIR Conference on Phase Change Materials and Slurries for Refrigeration and Air Conditioning, France.

148. Mehling H, Cabeza LF, Yamaha M. Phase change materials: application fundamentals. In: Pakso HÖ, ed. Thermal Energy Storage for Sustainable Energy Consumption. Springer 2007.

149. Memon SA. Phase change materials integrated in building walls: a state of the art review. Renew Sustain Energy Rev. 2014;31:870–906.

150. Midtdal K, Jelle BP. Self-cleaning glazing products: a state-of-the-art review and future research pathways. Sol Energy Mater Sol Cells. 2013;109:126–141.

151. Mortimer RJ, Rosseinsky DR, Monk PMS, eds. Electrochromic Materials and Devices. Wiley-VCH 2015.

152. Mosaffa AH, Ferreira CAI, Talati F, Rosen MA. Thermal performance of a multiple PCM thermal storage unit for free cooling. Energy Convers Manag. 2013;67:1–7.

153. Ng S, Sandberg LIC, Jelle BP. Insulating and strength properties of an aerogel-incorporated mortar based on UHPC formulations. Key Eng Mater. 2015a;629-630:43–48.

154. Ng S, Jelle BP, Sandberg LIC, Gao T, Wallevik ÓH. Experimental investigations of aerogel-incorporated ultra-high performance concrete. Constr Build Mater. 2015b;77:307–316.

155. Ng S, Jelle BP, Zhen Y, Wallevik ÓH. Effect of storage and curing conditions at elevated temperatures on aerogel-incorporated mortar samples based on UHPC recipe. Constr Build Mater. 2016;106:640–649.

156. Nguyen Q, Ngo T, Mendis P. A review on fire protection for phase change materials in building applications. In: Samali Attard, Song, eds. Materials to Structures: Advancement through Innovation. Taylor & Francis Group 2013.

157. Osterman E, Tyagi VV, Butala V, Rahim NA, Stritih U. Review of PCM based cooling technologies for buildings. Energy Build. 2012;49:37–49.

158. Özonur Y, Mazman M, Paksoy HÖ, Evliya H. Microencapsulation of coco fatty acid mixtures for thermal energy storage with phase change material. Int J Energy Res. 2006;30:741–749.

159. Pacheco-Torgal F. Indoor radon: an overview on a perennial problem. Build Environ. 2012;58:270–277.

160. Pacheco-Torgal F. Eco-efficient construction and building materials research under the EU framework programme Horizon 2020. Constr Build Mater. 2014;51:151–162.

161. Pacheco-Torgal F, Jalali S. Nanotechnology: advantages and drawbacks in the field of construction and building materials. Constr Build Mater. 2011;25:582–590.

162. Pasupathy A, Velraj R. Effect of double layer phase change material in building roof for year round thermal management. Energy Build. 2008;40:193–203.

163. Pomianowski M, Heiselberg P, Zhang Y. Review of thermal energy storage technologies based on PCM application in buildings. Energy Build. 2013;67:56–69.

164. Pomianowski M, Heiselberg P, Jensen RL, Cheng R, Zhang Y. A new experimental method to determine specific heat capacity of inhomogeneous concrete material with incorporated microencapsulated-PCM. Cem Concr Res. 2014;55:22–34.

165. Rao Z, Wang S, Peng F. Molecular dynamics simulations of nano-encapsulated and nanoparticle-enhanced thermal energy storage phase change materials. Int J Heat Mass Transf. 2013;66:575–584.

166. Rathod MK, Banerjee J. Thermal stability of phase change materials used in latent heat energy storage systems: a review. Renew Sustain Energy Rev. 2013;18:246–258.

167. RGEES, 2014a. Phase change material PIPE. Available from: <http://rgees.com/products_pcm-pipe.php> (accessed 07.01.2014).

168. RGEES, 2014b. Phase change material ball. Available from: <http://rgees.com/products_pcm-ball.php> (accessed 07.01.2014).

169. Ribeiro AS, Mortimer RJ. Electrochromic polymers. Encyclopedia of Polymer Science and Technology Wiley 2015.

170. Rodriguez-Ubinas E, Arranz BA, Sánchez SV, González FJN. Influence of the use of PCM drywall and the fenestration in building retrofitting. Energy Build. 2013;65:464–476.

171. Royon L, Karim L, Bontemps A. Thermal energy storage and release of a new component with PCM for integration in floors for thermal management of buildings. Energy Build. 2013;63:29–35.

172. Saihi D, Vroman I, Giraud S, Bourbigot S. Microencapsulation of ammonium phosphate with a polyurethane shell Part I: Coacervation technique. React Funct Polym. 2006a;64:127–138.

173. Saihi D, Vroman I, Giraud S, Bourbigot S. Microencapsulation of ammonium phosphate with a polyurethane shell Part II Interfacial polymerization technique. React Funct Polym. 2006b;66:1118–1125.

174. Salca, 2014. What’s K-Block. Available from: <http://www.salcabv.nl/index.asp?CategorieID=5&Taal=EN> (accessed 07.01.2014).

175. Sandberg LIC, Gao T, Jelle BP, Gustavsen A. Synthesis of hollow silica nanospheres by sacrificial polystyrene templates for thermal insulation applications. Adv Mater Sci Eng. 2013; 6 pages, Article ID 483651, 2013.

176. Sari A, Kaygusuz K. Some fatty acids used for latent heat storage: thermal stability and corrosion of metals with respect to thermal cycling. Renew Energy. 2003;28:939–948.

177. Sari A, Sari H, Önal A. Thermal properties and thermal reliability of eutectic mixtures of some fatty acids as latent heat storage materials. Energy Convers Manag. 2004;45:365–376.

178. Sarier N, Onder E. Organic phase change materials and their textile applications: an overview”. Thermochim Acta. 2012;540:7–60.

179. Schossig P, Henning H-M, Gschwander S, Haussmann T. Micro-encapsulated phase-change materials integrated into construction matrials. Sol Energy Mater Sol Cells. 2005;89:297–306.

180. Schröder J, Gawron K. Latent heat storage. Energy Res. 1981;5:103–109.

181. Schutzius TM, Jung S, Maitra T, et al. Physics of icing and rational design of surfaces with extraordinary icephobicity. Langmuir. 2015;31:4807–4821.

182. Sedeh MM, Khodadadi JM. Thermal conductivity improvement of phase change materials/graphite foam composites. Carbon. 2013;60:117–128.

183. Seppänen OA, Fisk WJ. Some quantitative relations between indoor environmental quality and work performance or health. ASHRAE Res J. 2006; February.

184. Schlanbusch RD, Jelle BP, Sandberg LIC, Fufa SM, Gao T. Integration of life cycle assessment in the design of hollow silica nanospheres for thermal insulation applications. Build Environ. 2014a;80:115–124.

185. Schlanbusch, R.D., Jelle, B.P., Sandberg, L.I.C., Fufa, S.M., Gao, T., 2014b. Life cycle assessment integrated in the design of a new nano insulation material. In: Proceedings of the 20th Annual International Sustainable Development Research Conference (ISDRC 2014), Trondheim, Norway, June, 18–20, pp. 325–335.

186. Sharma A, Sharma SD, Buddhi D. Accelerated thermal cycle test of acetamide, stearic acid and paraffin wax for solar thermal latent heat storage applications. Energy Convers Manag. 2002;43:1923–1930.

187. Sharma A, Tyagi VV, Chen CR, Buddhi D. Review on thermal energy storage with phase change materials and applications. Renew Sustain Energy Rev. 2009;13:318–345.

188. Shilei L, Neng Z, Guohui F. Impact of phase change wall room on indoor thermal environment in winter. Energy Build. 2006a;38:18–24.

189. Shilei L, Neng Z, Guohui F. Eutectic mixtures of capric acid and lauric acid applied in building wallboards for heat energy storage. Energy Build. 2006b;38:708–711.

190. Shrestha, S., Miller, W., Stovall, T., Desjarlais, A., Childs, K., Porter, W., et al., 2011. Modeling PCM-enhanced insulation system and benchmarking energyplus against controlled field data. In: Proceedings of the 12th Conference of International Building Performance Simulation Association, Sydney, November, 14–16.

191. Shukla A, Buddhi D, Sawhney RL. Thermal cycling test of few selected inorganic and organic phase change materials. Renew Energy. 2008;33:2606–2614.

192. Shukla N, Fallahi A, Kosny J. Performance characterization of PCM impregnated gypsum board for building applications. Energy Procedia. 2012;30:370–379.

193. Silva T, Vicente R, Rodrigues F. Literature review on the use of phase change materials in glazing and shading solutions. Renew Sustain Energy Rev. 2016;53:515–535.

194. Soares N, Costa JJ, Gaspar AR, Santos P. Review of passive PCM latent heat thermal energy storage systems towards buildings’ energy efficiency. Energy Build. 2013;59:82–103.

195. Solé A, Miró L, Barreneche C, Martorell I, Cabeza LF. Review of the T-history method to determine thermophysical properties of phase change materials (PCM). Renew Sustain Energy Rev. 2013;26:425–436.

196. Song JL, Guo QG, Zhong YJ, et al. Thermophysical properties of high-density graphite foams and their paraffin composites. New Carbon Mater. 2012;27:27–34.

197. Su W, Darkwa J, Kokogiannakis G. Review of solid–liquid phase change materials and their encapsulation technologies. Renew Sustain Energy Rev. 2015;48:373–391.

198. Sun Y, Wang S, Xiao F, Gao D. Peak load shifting control using different cold thermal energy storage facilities in commercial buildings: a review. Energy Convers Manag. 2013;71:101–114.

199. Sveipe E, Jelle BP, Wegger E, et al. Improving thermal insulation of timber frame walls by retrofitting with vacuum insulation panels—experimental and theoretical investigations. J Build Phys. 2011;35:168–188.

200. Tate, 2013. EcoCore. <http://www.tateinc.com/products/ecocore.aspx> (accessed 18.12.2013).

201. Tatsidjodoung P, Pierrès NL, Luo L. A review of potential materials for thermal energy storage in building applications. Renew Sustain Energy Rev. 2013;18:327–349.

202. Tenpierik MJ. Vacuum insulation panels applied in building constructions (VIP ABC) PhD thesis Technische Universiteit Delft 2009.

203. Thambidurai M, Panchabikesan K, Mohan, N K, Ramalingam V. Review on phase change material based free cooling of buildings—the way toward sustainability. J Energy Storage. 2015;4:74–88.

204. Tyagi VV, Kaushik SC, Tyagi SK, Akiyama T. Development of phase change materials based microencapsulated technology for buildings: a review. Renew Sustain Energy Rev. 2011;15:1373–1391.

205. Van Den Bergh S, Hart R, Jelle BP, Gustavsen A. Window spacers and edge seals in insulating glass units: a state-of-the-art review and future perspectives. Energy Build. 2013;58:263–280.

206. Waqas A, Din ZU. Phase change material (PCM) storage for free cooling of buildings—a review. Renew Sustain Energy Rev. 2013;18:607–625.

207. Wegger E, Jelle BP, Sveipe E, et al. Aging effects on thermal properties and service life of vacuum insulation panels. J Build Phys. 2011;35:128–167.

208. Weinläder H, Beck A, Fricke J. PCM-facade-panel for daylighting and room heating. Solar Energy. 2005;78:177–186.

209. Xiao J, Chaudhuri S. Design of anti-icing coatings using supercooled droplets as nano-to-microscale probes. Langmuir. 2012;28:4434–4446.

210. Xu X, Zhang Y, Ling K, Di H, Yang R. Modeling and simulation on thermal performance of shape-stabilized phase change material floor used in passive solar buildings. Energy Build. 2005;37:1084–1091.

211. Yang C, Fischer L, Maranda S, Worlitschek J. Rigid polyurethane foams incorporated with phase change materials: a state-of-the-art review and future research pathways. Energy Build. 2015;87:25–36.

212. Yu S, Jeong SG, Chung O, Kim S. Bio-based PCM/carbon nanomaterials composites with enhanced thermal conductivity. Sol Energy Mater Sol Cells. 2014;120:549–554.

213. Yuan Y, Zhang N, Tao W, Cao X, He Y. Fatty acids as phase change materials: a review. Renew Sustain Energy Rev. 2014;29:482–498.

214. Zalba B, Marín JM, Cabeza LF, Mehling H. Review on thermal energy storage with phase change: materials, heat transfer analysis and applications. Appl Thermal Eng. 2003;23:251–283.

215. Zalba B, Marín JM, Cabeza LF, Mehling H. Free-cooling of buildings with phase change materials. Int J Refrigeration. 2004;27:839–849.

216. Zhang Y, Jiang Y. A simple method, the T-history method, of determining the heat of fusion, specific heat and thermal conductivity of phase change-materials. Meas Sci Technol. 1999;10:201–205.

217. Zhang P, Lv FY. A review of the recent advances in superhydrophobic surfaces and the emerging energy-related applications. Energy. 2015;82:1068–1087.

218. Zhang H, Xu Q, Zhao Z, et al. Preparation and thermal performance of gypsum boards incorporated with microencapsulated phase change materials for thermal regulation. Sol Energy Mater Sol Cells. 2012;102:93–102.

219. Zhang P, Xiao X, Ma ZW. A review of the composite phase change materials: fabrication, characterization, mathematical modeling and application to performance enhancement. Appl Energy. 2016;165:472–510.

220. Zhao CY, Zhang GH. Review on microencapsulated phase change materials (MEPCMs): fabrication, characterization and applications. Renew Sustain Energy Rev. 2011;15:3813–3832.

221. Zhao CY, Lu W, Tian Y. Heat transfer enhancement for thermal energy storage using metal foams embedded within phase change materials (PCMs). Solar Energy. 2010;84:1402–1412.

222. Zhong Y, Guo Q, Li S, Shi J, Liu L. Heat transfer enhancement of paraffin wax using graphite foam for thermal energy storage. Sol Energy Mater Sol Cells. 2010;94:1011–1014.

223. Zhou D, Zhao CY, Tian Y. Review on thermal energy storage with phase change materials (PCMs) in building applications. Appl Energy. 2012;92:593–605.

224. Zhu N, Ma Z, Wang S. Dynamic characteristics and energy performance of buildings using phase change materials: a review. Energy Convers Manag. 2009;50:3169–3181.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.191.46.36