Skip to content
BY 4.0 license Open Access Published by De Gruyter September 21, 2020

Phase change materials for building construction: An overview of nano-/micro-encapsulation

  • Amende Sivanathan , Qingqing Dou , Yuxuan Wang , Yunfeng Li , Jorge Corker , Yonghui Zhou and Mizi Fan EMAIL logo
From the journal Nanotechnology Reviews

Abstract

Buildings contribute to 40% of total global energy consumption, which is responsible to 38% of greenhouse gas emissions. It is critical to enhance the energy efficiency of buildings to mitigate global warming. In the last decade, advances in thermal energy storage (TES) techniques using phase change material (PCM) have gained much attention among researchers, mainly to reduce energy consumption and to promote the use of renewable energy sources such as solar energy. PCM technology is one of the most promising technologies available for the development of high performance and energy-efficient buildings and, therefore, considered as one of the most effective and on-going fields of research. The main limitation of PCM is its leakage problem which limits its potential use in building construction and other applications such as TES and textiles, which can be overcome by employing nano-/micro-encapsulation technologies. This paper comprehensively overviews the nano-/micro-encapsulation technologies, which are mainly classified into three categories including physical, physiochemical and chemical methods, and the properties of microcapsules prepared. Among all encapsulation technologies available, the chemical method is commonly used since it offers the best technological approach in terms of encapsulation efficiency and better structural integrity of core material. There is a need to develop a method for the synthesis of nano-encapsulated PCMs to achieve enhanced structural stability and better fracture resistance and, thus, longer service life. The accumulated database of properties/performance of PCMs and synthesised nano-/micro-capsules from various techniques presented in the paper should serve as the most useful information for the production of nano-/micro-capsules with desirable characteristics for building construction application and further innovation of PCM technology.

1 Introduction

Phase change materials (PCMs) are a group of functional materials that support the same purpose as a function of temperature with intrinsic capability of absorbing, storing and releasing thermal energy in form of latent heat known as enthalpy of fusion [1,2,3], during phase transition cycles at their operating temperatures under isothermal conditions. In general, PCMs are distinguished according to the phase change they undergo at a particular temperature level and are classified based on their physical transformation of absorbing/storing or releasing thermal energy capabilities, as result of undergoing a particular phase transformation from one state to another (e.g. solid to liquid or liquid to gas) at their operating temperatures [4]. It is evident from the critical examination that PCMs can enhance the performance as well as the safety, with respect to their premium energy storage. Due to the numerous unique advantages they possess including high energy density, abundance in nature, wide temperature range and low cost, they are often concerned for building materials, electronic devices, lithium-ion-battery, and in solar energy systems to improve the heat storage of the water tanks to enhance the solar energy utilisation efficiency [5,6]. Overall, PCM application because of their aforementioned intrinsic characteristics ranges from the hand warmer to green buildings; however, for most applications, PCMs with a narrow phase change temperature range with a corresponding high latent heat value, along with robust chemical stability, are preferred because they can store and release higher amounts of energy cycles. Availability of PCMs in a wide range of melting points [7] is the main reason why they are extensively utilised in several different real-world applications including the following: (1) developing smart thermal microgrids; (2) portable thermal batteries; (3) indoor thermal management systems; (4) thermoregulating textiles, (5) warm supplies thermal protection; (6) in solar-driven cookers, (7) in solar heating systems; (8) water heaters; (9) refrigerators; (10) air-conditioning; (11) for cooling; (12) for the enhancement of thermal comfort in buildings; (13) thermal performance of building materials; (14) for the fabrication of energy-saving equipment, (15) in healthcare; and (16) food preservation [8,9,10]. Overall, they are highly beneficial in any application that relies on controlled and efficient thermal energy storage (TES) and releases, commonly achieved by exploiting their melting and crystallisation behaviour. It is critical that PCM selected for a designated practical purpose has a phase change temperature (e.g. operating/working temperature) within the temperature range of the given application; otherwise, they will behave like conventional sensible heat storage materials [11]. The suitable working temperature ranges of textile and building applications are 18–35°C and 19–26°C [12].

PCMs are considered to play a vital role since a few decades in efficient latent heat storage applications, as advanced latent heat thermal energy storage (LHTES) materials [13], for their incorporation in TES systems enable adequate control over the operating parameters of the system through the utilisation of the exothermic and endothermic phase transitions, which allows thermal energy to be stored more effectively [14]. Theoretically, they can store 5–14 times more thermal energy per unit volume than conventional sensible heat storage materials such as rock, masonry, and water [15], and their high storage density enables the attainment of a compact energy storage system, operating under isothermal conditions [16]. Recently, the principle of PCM is being exploited and developed in building energy conservation systems and in solar heating, mainly to enhance their energy efficiency [17,18], since it is considered as the most practical solution available to reduce the volume of fossil fuel consumed. Primarily, because the building sector is one of the largest energy-consuming sectors; so, it is critical to advance the building energy performance; the increased demand of fossil fuel is posing critical environmental problems such as rapid climate change, along with threatening global economy due to the increasing costs of fossil materials and finite resources. Employing PCM technology in buildings will lower the heating and cooling loads by reducing the heat transferred through the building envelope, so that the indoor temperature will be maintained within the thermal comfort range of occupants by reducing temperature swings, and consequently, the total energy consumption of building will be reduced significantly by load reduction or shifting [2]. The thermal performance of PCMs integrated in buildings depends entirely upon the melting temperature, thermal conductivity and TES density. PCMs with rapid melting and crystallisation/solidification points are considered as suitable candidates for TES applications [19]. Designing a heat storage system is a complex task since it is not restricted to the thermal performance of the system only, but also involves taking into consideration the cost, safety and sustainability of materials used and processes employed [20].

In the past, various PCM integration methods have been reported such as direct incorporation, immersion, shape-stabilization and encapsulation, e.g. nano-, micro- and macro-encapsulation [21]; however, latter approach has not proven to be practical in the past [22]. Direct incorporation of PCM can be done by the means of the imbibing technique (e.g. direct incorporation), in which construction materials like gypsum boards [23], concrete blocks, or bricks [24] get immerged into PCM baths, which results in PCM being adsorbed onto the pores of the construction materials; despite being practical, final PCM-filled materials obtained has faced serious phase segregation problems due to PCM leakage, in addition to difficulty in moisture transfer across the enclosure [25]. Also, PCM direct integration into building material requires special latent heat devices or heat exchange surfaces, which leads to increased associated costs and increased thermal resistance between the PCM and the environment; therefore, not many building materials will be able to incorporate PCMs directly.

Like any other technologies, PCMs are also facing challenges. First, an ideal PCM is expected to meet a number of criteria with respect to the thermophysical properties including high enthalpy per unit volume, suitable phase change temperature, no supercooling, long-term chemical stability and high thermal conductivity. Unfortunately, aforementioned merits are rarely possessed fully by most PCMs, which means these shortcomings may result in degrading the performance of the system used for TES and thereby hinder their final applications [26]. Also, construction materials capable of incorporating PCMs are relatively limited, since PCMs have poor compatibility with most buildings materials. Well-known drawbacks of PCMs includes the following: (1) leakage problem that occurs during the phase transition process and (2) their inherently low thermal conductivity (e.g. poor working efficiency). Leakage problem can cause serious damages, such as contamination or device failure if PCM is used in packaging or electronic devices [1,10], and low thermal conductivity means rate at which PCM absorbs/stores and releases energy will be extremely slow making them unsuitable for most applications. The leakage problem is most serious one since it undesirably suppresses the potential practical application of PCMs in buildings and other applications significantly; however, this can be overcome by employing either the nano/micro-encapsulation or shape stabilization technology. Nano-/micro-encapsulation technology involves encapsulating PCM with a continuous polymeric material, whereas in the shape-stabilization technology, the chosen PCM gets embedded into a supporting material. However, latter technology is considered as the most effective strategy to overcome the aforementioned issues, and it also encourages industrial scale production of powder or granulates of PCMs composites, which surely widens the potential scope of PCMs, especially organic PCMs in real applications [10]. Although shape-stabilisation technology is a highly feasible method, in terms of economy and practicality, it is impossible to substitute nano-/micro-encapsulation technology with it completely in certain applications, e.g. for thermal regulation in buildings and textiles, the use of PCM microcapsules is required, while PCM composite resulting from shape-stabilisation technology is undesirable. Nano-/micro-encapsulation technology usually encapsulates organic PCMs including fatty acids, paraffin within polymers because they exhibit numerous benefits such as appropriate calorific value, non-corrosion and non-supercooling properties [27]; however, they display major drawbacks such as low thermal conductivity and flammability, which tends to limit their real-world applications. On the other hand, inorganic PCMs do have number of benefits such as high energy density, relatively high thermal conductivity and low cost, but their applications are hindered significantly due to the following two serious drawbacks they exhibit: (1) phase separation/segregation and (2) sub-cooling/supercooling behaviours [28].

2 Classification of PCMs

PCMs used for LHTES systems can be classified into four states according to the phase change mechanism they are subjected to and the phase transition temperature: (1) solid–solid (S–S), (2) solid–liquid (S–L), (3) solid–gas (S–G) and (4) liquid–gas (L–G) (Figure 1). However, for practical purposes, solid–liquid PCMs are commonly used because other varieties such as solid–gas and liquid–gas are not compatible with building materials [29] due to some technical limitations including the following: (1) large volume changes during the phase transition process and (2) existence of high pressure of the gas phase in the system [30,31]; therefore, it will not be discussed further. Despite the fact that a wide variety of PCMs with different melting point ranges are available on the market, the most commonly used type of solid–liquid PCMs are categorised into three types: (1) organic, (2) inorganic and (3) eutectic mixtures. Literature reports that solid–solid PCMs have one major advantage over solid–liquid PCMs [32], i.e. they require no encapsulation since no fluid (i.e. gas or liquid) gets generated during the phase transition process; so, the overall cost will be reduced because nano-/micro-encapsulation technology will not be required. Furthermore, it allows direct incorporation of PCM in building materials and/or building components. Also, solid–solid PCMs are believed to have good compatibility with building materials; however, till date no study has investigated the compatibility between the synthesised solid–solid PCMs and building materials. In general, a PCM undergoes phase change as a result of getting activated by a stimulus such as heat, light or pressure/strain. PCMs have different with some of them having multiple of the mentioned stimuli for their activation, and their phase change occurs within their specific boundary conditions such as changes in temperature, light, concentration of solvents and pH [33]. Several factors tend to influence their effectiveness and applications such as good thermal properties depend entirely upon suitable phase transition temperature, high latent heat of transitions per unit weight, high thermal conductivity and large specific heat capacity. In addition, PCMs are also expected to possess other characteristics such as a favourable phase equilibrium and high density. Critical kinetic properties that PCMs should exhibit are little or no sub-cooling effect during freezing, possess sufficient crystallization rate and solidification and melting should occur at the same temperature without phase segregation. Overall, PCMs should have the following desirable characteristics: (1) good chemical properties to ensure completion of reversible freezing/melting cycle, (2) long-term chemical stability, (3) good compatibility with other materials, (4) undergo no degradation after being subjected to long-term thermal cycles, (5) non-corrosive, (6) non-toxic, and (7) no fire hazard and should contain non-explosive compounds only [34]. Till date, literature has majorly reported microencapsulation of commercially available organic PCMs such as paraffin due to their non-corrosive and hydrophobic nature, which are main desirable characteristics that PCMs are expected to have, to avoid deterioration of thermal and structural strength of encapsulated PCM after being subjected to large number of thermal cycles. Hydrophilic nature of PCM will result in adsorption of water molecule, and this will influence the properties of PCM greatly and thus its performance. Therefore, encapsulation of organic PCMs is preferred over other PCMs such as salt hydrates [35]. Solid–liquid PCMs are commonly utilised as efficient storage materials currently since they are capable of absorbing, storing or releasing large amounts of heat energy, almost isothermally by undergoing phase change, while having relatively high heat capacity and storage, and they experience relatively smaller volume changes during the phase transition process when compared with the liquid–gas PCMs [1], and they have higher latent heat storage capacities than the solid–solid PCMs. The PCM to be used for an application in various fields, such as construction, refrigerators, air-conditioning, textiles, food preparation, astronautics and waste heat recovery, is determined based on the temperature at which phase change occurs, better known as its operating temperature [36]. Vast number of solid–liquid PCMs have been prepared in the past to deduce a solution that provides a highly effective storage system at low costs. However, this has not been possible yet, since solid–liquid PCMs have poor shape-stability, making it difficult for them to retain their original solid state during the phase transition process, so to encourage its potential application, the encapsulation technology must be employed, which results in increasing the overall costs greatly. In addition, solid–liquid PCMs face other operational problems such as supercooling and corrosion [37]. Nevertheless, PCMs have their superior properties for one over the others as summarised in Table 1. These advantages and disadvantages of different PCMs lead to various processing, formulations and end-user requirements.

Figure 1 
               Classification of PCMs.
Figure 1

Classification of PCMs.

Table 1

Typical advantages and disadvantages of organic, inorganic and eutectic mixtures PCMs

PCM type Advantages Disadvantages Ref.
Organic
  • No supercooling

  • No phase segregation during the crystallisation process

  • High latent heat of energy storage

  • High resistance to corrosion

  • No toxicity

  • Smaller volume change

  • High compatibility with construction materials

  • Can be incorporated directly

  • Can be impregnated into materials

  • Good chemical stability and inert i.e. unreactive

  • High TES

  • High range of melting temperatures

  • Low vapor pressure during phase change

  • Good thermal performance even after subjection to prolonged thermal cycling, i.e. good thermal behaviour

  • Excellent freezing and melting properties

  • Good thermal stability

  • Adjustable transition zone

  • Good nucleation rate

  • Recyclable

  • Self-nucleation

  • Congruent-melting

  • Low thermal conductivity (e.g. typically around 0.2 W/m K for paraffins)

  • Large volume change (especially paraffins)

  • Expensive, i.e. non-paraffins are around three times more expensive than paraffins

  • High flammability

  • Low phase change enthalpy

  • Low volumetric storage density and melting point

  • High volatility

[11,31,38,39]
Inorganic
  • High TES density

  • High thermal conductivity

  • Less expensive when compared with organic PCMs.

  • Non-flammable

  • High phase change enthalpy and sharp phase change

  • Small volume change

  • High latent heat of fusion

  • Recyclable

  • Prone to corrosion, i.e. corrosive to most metals

  • Some degree of toxicity

  • Requires container and support

  • Supercooling

  • Phase segregation

  • Low thermal performance after subjection to prolonged thermal cycling, i.e. poor thermal behaviour

  • Cannot be incorporated directly

  • Decomposition

  • Moderately compatible with construction materials, i.e. not compatible with all building materials

  • Incongruent melting and dehydration during thermal cycling

  • Poor nucleating properties – nucleating and thickening agents are required to their applications

  • Low-temperature range

[11,31,38]
Eutectic mixtures
  • Sharp melting points

  • High volumetric storage density

  • No phase segregation

  • No congruent phase change

  • Some of them suffer from super-cooling

  • Low total latent heat capacity

  • Strong odours

  • Limited data are available in the literature on their thermophysical properties

  • High cost

[11,31,38]

3 Nano-/micro-encapsulation of PCM

3.1 Encapsulation technologies

Encapsulation technology was first introduced and developed in 1940s and 1950s by Barett K. Green [35,40], when it was called microencapsulation, and since then it has been developed further, researched and applied in various sectors including chemistry, biology and medicine due to the remarkable advantages it exhibits [1,41]. Overall, it is possible to develop value-added products using this microencapsulation technique, and it can be defined as the process by which one or more compounds in form of individual solid particle or liquid droplets (e.g. the core) get surrounded by a protective layer, by the means of a continuous polymeric film, composed of one or more materials (e.g. the shell), to produce a core–shell structure; or by which the internal core material gets dispersed throughout the protective material matrix, mainly to isolate, and thereby protect the core substance (PCM) from the surrounding, which is composed of various external environmental factors such as light, high oxygen concentration, heat, moisture and also to prevent volatile compounds from evaporating, to mask unpleasant odours and tastes or to release the internal material at controllable speeds [1,42]. Microencapsulation of core material ensures significant reduction in the reactivity of the core material, promotes controlled release of the core materials and, in addition, eases the handling, application and storage of the core material. However, its function depends highly upon its intended final application. The morphology of microcapsule is influenced by the core material as well as the processing technique applied [43]. Figure 2 shows the appearance of a typical microcapsule. In general, there are three main requirements that have to be established for successful encapsulation of PCMs: (1) the formation of shell surrounding the PCM core completely, (2) formed shell must prevent the leakage of melted PCM and (3) no incorporation of impurities in the core/shell system. Therefore, nano-/micro-capsule synthesised needs to be capable of withstanding the applied mechanical and thermal stresses. Capsules resulting from microencapsulation technique is termed according to their size and are namely nanocapsules, microcapsules or macrocapsules. The main difference between nano-, micro- and macrocapsules lies in the size and shape of capsules produced, which are influenced greatly by the synthesis approach applied and the type of shell material used. If capsules are composed of a diameter in the range 0 to 1,000 nm (e.g. in the colloidal range), they are classified as nanoparticles, better known as nanocapsules, nanoparticles or nanospheres [44,45,46]. If capsule contains a diameter less than 1 mm or 1 cm, they are called microcapsules, where capsules of diameters greater than 1 mm or 1 cm are classified as a macrocapsules [28]. Figure 3 illustrates how capsules are classified according to their size. In general, microencapsulation and nanoencapsulation processing techniques tend to be more complicated, when compared with the macroencapsulation processing, since the capsule size is much smaller (e.g. less than 1 mm); however, reducing the capsule size is more advantageous since such particles have higher structural stability and better fracture resistance [47]. Macroencapsulation is usually conducted by filling the prepared shells with liquid PCMs, and the resulting macrocapsules are produced by encapsulating PCM in pouches, shells, tubes, thin plates and spheres [28]. Mentioned encapsulation techniques are reported to be advantageous because they allow for more flexible choices of PCM and shell material selection, according to the practical requirement. But to achieve optimum performance, correct selection of PCM and shell materials needs to be made, usually accomplished by following a standard selection procedure. The major drawback of microencapsulated PCMs is their inherently low thermal conductivity because the shell material (e.g. polymer) is of relatively low thermal conductivity [9].

Figure 2 
                  Appearance of a typical microcapsule containing a PCM core.
Figure 2

Appearance of a typical microcapsule containing a PCM core.

Figure 3 
                  Classification of capsules according to their size.
Figure 3

Classification of capsules according to their size.

3.2 Encapsulation methodologies

There are numerous types of physical and chemical methods developed in the past for the synthesis of microcapsules; however, chemical and physical methods were commonly used for the microencapsulation of PCMs. The latter technique can be subdivided into two types, i.e. physio-chemical and physio-mechanical [48]. The commonly used physical methods are pan-coating, air-suspension coating, spray drying, centrifugal extrusion and vibrational nozzle, while commonly used chemical methods are in situ polymerisation, interfacial polymerisation, suspension polymerisation and emulsion polymerisation [49]. The type of microencapsulation technique to be used is deduced by various factors such as (1) the type of shell material used (e.g. chemical composition of material), (2) the nature of the core material to be encapsulated (e.g. hydrophobic or hydrophilic) and (3) the release mechanisms to be encountered for the intended application. Overall, it depends entirely upon the physical and chemical properties of the materials used [8,49]. Selection of encapsulation technique, shell material and core material is made based on the final application of the product. Therefore, factors such as physical and chemical stability, concentration, required particle size and manufacturing cost should be considered because they dictate the final performance of the synthesised capsules [48]. Also, each microcapsule to be produced has to meet particular demands concerning the properties of the capsule, i.e. particle size/particle size distribution, thickness of shell and mechanical characteristics [50]. The employed technique should be controlled to prevent fracture of synthesised capsules caused by collision, and the strength of shell formed should be adequate to resist stress conditions, such as volumetric changes generated during the phase transition process [35]. Figure 4 shows different microencapsulation approaches available. It is impossible to produce microcapsules of average diameter less than 100 µm by employing the physical method [51], and therefore, for the synthesis of nanocapsules, chemical methods should be employed [29].

Figure 4 
                  Different microencapsulation technologies.
Figure 4

Different microencapsulation technologies.

3.2.1 Shell materials

Characteristics and performance of microcapsules synthesised by the employment of various approaches are influenced greatly by the chemical and physical properties of the shell material used, and therefore, the selection of shell material is very important. There are in total three different types of shell materials such as organic, inorganic and organic–inorganic hybrid shell.

3.2.1.1 Organic shell material

Examples of organic shells reported in the past are mainly melamine-formaldehyde (MF) resin, urea-formaldehyde (UF) resin, poly(urea-urethane), polyurea (PU) and acrylic resins, and these types of shell materials are known to offer structural stability to the microcapsules, since they have excellent durability during the repetitive phase change process; however, they exhibit poor chemical and thermal stabilities. Out of all organic shell materials available, acrylic resins that are copolymers of methacrylate are commonly used for the microencapsulation of PCMs because they consist of several advantages such as non-toxicity, easy handling, high mechanical strength and chemical resistance, along with good optimal clarity. Resulting microcapsules can be used in thermal management and conservation applications in the solar energy technology because poly(methyl methacrylate) (PMMA) shell material exhibits exceptional sunlight stability [10].

3.2.1.2 Inorganic shell material

Commonly used inorganic shell materials are silica, titania, calcium carbonate, polystyrene (PS) and zinc oxide. Literature reports that utilisation of inorganic shell material is better than the organic ones because they exhibit numerous desirable characteristics such as excellent thermal stability, relatively high thermal conductivity, better mechanical durability, chemical inertness, and non-toxicity. Therefore, synthesised microcapsules can be used for thermal management of buildings and in automotive applications. Silica (SiO2) is one of the extensively studied inorganic material for the microencapsulation of organic PCMs such as paraffin [10] because it is susceptible to surface modification with different functional groups [52]. Compared with organic shell materials that possess drawbacks including toxicity, flammability, poor heat transfer performance and poor thermal stability, inorganic shell materials are more advantageous and, therefore, have gained more attention in the recent times [53]. Encapsulating flammable organic PCM like paraffin wax within an inorganic shell material tends to reduce its flammability [35].

3.2.1.3 Organic–inorganic hybrid shell material

They exhibit outstanding characteristics when compared with both organic and inorganic shell materials such as superior thermal conductivity, chemical stability, and mechanical robustness, and hence, aforementioned properties are achieved by incorporating inorganic additives such as silver nanoparticles, iron nanoparticles or silicon nitride in the organic shell. One of the common drawbacks of employing this approach is the superficial attachment of inorganic additives on the microcapsules, and in addition, they are observed to detach from the surface during the repetitive phase transition process. However, it is possible to enhance this approach by employing chemical hybridisation of organic and inorganic shell materials, which usually results in improving the mechanical strength of shell material significantly, and its thermal performance [10].

3.2.2 Physical encapsulation

3.2.2.1 Spray drying process

Spray drying process is a simple, convenient, reproducible and scalable drying technology with significant encapsulation efficiency, first developed in the 1930s, and since then it has attracted many due to the advantages it exhibits such as comparatively low production cost, uses less energy, and it is ideal for temperature sensitive materials (e.g. food and biological items) [54,55]. It is one of the oldest encapsulation methods commonly used due to the feasible processing steps it consists of [56]. The simple processing steps of this method includes the following: (1) heating the drying gas, (2) generation of droplets, (3) drying of droplets formed and (4) collection of isolated particles produced (see Figure 5), and they are comparable to the “one stage drying operation.” It allows various liquids to be converted into solid particles with adjustable size, distribution, shape, porosity, density, and chemical composition [55], and with this process, it possible to produce a wide range of microcapsules at good yield, and the synthesised microcapsules are reported to have good encapsulation efficiency. Furthermore, it covers a range of feedstock and product applications such as core material used can be in solution, suspension, slurry or molten form. This method is well established in the food industry [57], pharmaceutical industry [55,58] and lately to microencapsulate PCMs [41]. However, it does have drawbacks such as the spray drying equipment is very expensive and bulky. Moreover, it produces microcapsules in powder form, which requires further processing (e.g. agglomeration), and the resulting product has very low overall thermal efficiency, mainly because a large volume of heated air gets passed at high velocity through the chamber without contacting the particles and, therefore, makes no direct contribution to drying. Conversely, this technique is limited by the number of shell materials existing, i.e. shell material should be soluble in water (e.g. hydrophilic) [48]. Hawlader et al. [41] fabricated microcapsules containing paraffin-wax as a core material with gelatine/acacia as shell materials by this spray drying method. The microencapsulation efficiency was observed to decrease with increasing core/shell ratio due to the presence of insufficient amount of shell material for complete encapsulation. Addition of cross-linking agent provided the amount that does not exceed the range of 6–8 mL, resulted in enhancing the microencapsulation efficiency greatly. The presence of crosslinking agent in excess (i.e. greater than 8 mL) caused alteration of the mechanism and inability to integrate into the network. Borreguero et al. [59] microencapsulated paraffin Rubitherm ®RT27 with and without carbon nanofibers (CNFs) within a low density polyethylene ethyl-vinyl-acetate shell. Results demonstrated that the particle size and shape of microcapsules collected from the collection vessel were better than the ones collected from the inside of the drying vessel. Hence, the microcapsules obtained from the inside of the drying chamber was bound together due to the melting of the polymer (e.g. they had agglomerated), whereas the ones collected from the collection vessel exhibited a more homogeneous and spherical shape with an average diameter of 3.9 µm and a paraffin content of 49%. The stiffness of resulting microcapsules was observed to have increased after incorporating CNFs, and thus, observations showed that CNFs were mainly found inside the microcapsules or had formed bridges between the microcapsules which led to the modification of the final particle size. Adding CNFs influenced the morphology of the microcapsules, and thereby, it promoted agglomeration. Thermal stability studies conducted on over 3,000 cycles indicated that the wax present between the synthesised microcapsules was melted and solidified in a reversible manner during the 30 years of continuous operation time, and the results indicated that these microcapsules exhibited greater thermal stability when compared with ones formed with PS. Fei et al. [60] microencapsulated n-octadecane (C18) within a titania shell via a rapid aerosol process with a hydrothermal post-treatment. Results showed that microcapsules with 80% C18 content were obtained with diameters in the range of 0.1–5.0 µm with two structures of dense or hollow structures, and the microcapsules exhibited an enthalpy in the range of 92–97 J/g.

Figure 5 
                        Schematic representation of the spray drying method employed for the synthesis of isolated microcapsules.
Figure 5

Schematic representation of the spray drying method employed for the synthesis of isolated microcapsules.

3.2.2.2 Pan-coating

In this pan-coating method, widely used in the pharmaceutical industry, small-coated capsules are formed by mixing dried coating material with the solid particles, followed by melting of the coating material to ensure that solid capsules get surrounded with the coating material, and finally, cooling to enable solidification. In an alternative approach, the coating material gets sprayed gradually on the core material, by the means of tumbling in a vessel, unlike the conventional pan-coating method, which involves mixing coating material at the very start of the encapsulation process; however, this alternative approach is considered to be better, since it has shorter processing time and reduced costs [8].

3.2.2.3 Air suspension coating

This technique provides enhanced controls and flexibility when compared with the pan coating technique because it operates by altering the times at which the core particles pass the coating zone. During this process, the solid particles get coated and dried while remaining in their suspended position, by the means of an air stream moving in the upwards direction. This technique is reported to be unsuitable for the encapsulation of PCMs; however, they are commonly applied in the pharmaceutical industry, food industry and the production of cosmetics [8].

3.2.2.4 Centrifugal extrusion

It is a process developed by Southwest Research Institute. In this process, the core material, in liquid form, passes through the inner tube and the coating material, which is usually immiscible with the core material, passes through an annular tube around it. Vibration of the head results in emerging the core and coating material from the orifice at the ends of the tube, then they get broken down into spherically shaped drops, caused by the tensile forces present on the surface. No previous studies have used this method to encapsulate PCM. These drops will get solidified in a bath by the means of heating or suitable chemical reactions [8].

3.2.3 Physic-chemical encapsulation

3.2.3.1 Ionic-gelation

This is a simple and easy method commonly used in the pharmaceutical industry, mainly in drug delivery systems and food industry for the synthesis of nanoparticles. In this method, a hydrophobic drug such as curcumin gets dissolved completely in the chosen solvent, followed by the addition of the polymer solution, under constant stirring condition. The encapsulation efficiency is influenced greatly by the degree of crosslinking that has occurred along the polymer [61]. Also, it is based on the ability of polyelectrolytes to crosslink by the application of either extrusion or emulsification/gelation, in the presence of multivalent counter ions such as Ca2+, Ba2+, and Al3+ to produce the hydrogels. For instance, the ionic gelation of alginate and calcium ion results in the formation of calcium alginate (CaAlg) microcapsules [8]. This method has not yet been employed for the encapsulation of PCMs.

3.2.3.2 Complex coacervation

This is a term commonly used in colloidal chemistry to indicate the associative phase separation process that gets induced as a result of modifying the environment of the media such as pH, ionic strength, temperature or solubility under controlled conditions (see Figure 6) and can be defined as a colloidal phenomenon that involves liquid–liquid phase separation of a single or a mixture of two polymers of opposite charges in aqueous solution, resulting from triggering caused by electrostatic interactions, hydrogen bonding, hydrophobic interactions and/or polarisation-induced attractive interactions. Chemical or enzymatic crosslinking agents commonly used includes glutaraldehyde or transglutaminase. The degree of interaction present between the biopolymers depend highly upon factors such as biopolymer type, e.g. molar mass, flexibility and charge, pH, ionic strength, concentration and ratio of biopolymers [62]. In this process, the phase rich in colloidal is called the coacervate phase, while the phase containing minor amounts of colloidal is called the equilibrium phase [63]. There are two types of coacervation processes: (1) simple and (2) complex; however, both processes will have the same microcapsule formation mechanism, with only exceptions being the polymer system involved in the reaction, and the mechanism of phase separation (e.g. phase separation process varies in both processes). In the simple coacervation method, a desolvation agent gets added during the phase separation process, whereas the complex coacervation method is more about the complexation taking place between the oppositely charged polymers and usually results in the synthesis of stable and small-sized particles. Hence, complex coacervation process initiates through the interaction between the oppositely charged polymers or macromolecules such as proteins or polyelectrolytes [64]. The initial step involves the preparation of the emulsion by dispersing the core material (oil) into an aqueous polymer solution. Addition of second polymer solution followed by the addition of salt or by altering the pH temperature or by dilution of the medium results in the deposition of the shell material onto the core particles. The microcapsules obtained are then stabilised by crosslinking, desolvation or thermal treatment, and nano-/microparticles are usually collected by centrifugation. This technique is inexpensive and makes use of environmentally friendly solvents only; however, major shortcoming is that it requires the use of large amount of solvents [61]. Figure 6 illustrates the processes involved in the complex coacervation method.

Figure 6 
                        Schematic representation of the fundamental steps involved in a complex coacervation process.
Figure 6

Schematic representation of the fundamental steps involved in a complex coacervation process.

Hawlader et al. [65] microencapsulated paraffin wax within a gelatin/gum acacia shell by the complex coacervation method. Optimising processing conditions was determined based on the response surface method. Results indicated that increasing the paraffin wax/shell material ratio will improve the paraffin encapsulation ratio, and thereby, increases the content of paraffin wax encapsulated. At higher encapsulation ratio of paraffin wax, microcapsules exhibiting lower hydrophilicity were obtained. Hawlader et al. [41] used the same method to fabricate microcapsules containing paraffin-wax as a core material with gelatine/acacia as shell materials. Onder et al. [66] microencapsulated different paraffin waxes, namely n-hexadecane, n-octadecane and n-nonadecane with natural and biodegradable polymers, namely gum Arabic–gelatin mixture. Deveci and Basal [67] microencapsulated n-eicosane within a chitosan (CHI)/silk fibroin (SF) shell. Results demonstrated successful production of spherical particles by the coacervation process, and it confirmed the possibility of using SF/CHI wall materials to encapsulate PCM. Jun-xia et al. [68] microencapsulated sweet orange oil within a soybean protein (SPI)/gum Arabic (GA) shell. Results showed that pH 4 was determined as the electrical equivalence point, and the optimum SPI/GA ratio was 1.0. Observations showed that high ionic strength was non-beneficial for this process; however, adding sucrose resulted in enhancing the microencapsulation efficiency and product yield significantly. Bayés-García et al. [69] microencapsulated Rubitherm®R27 with two different shell material compositions namely sterilised gelatine/arabic gum (AG) and agar–agar (AA)/AG by the complex coacervation method and reported encapsulation ratios of 49 and 48%, respectively. In general, caprylic acid is considered as the most promising fatty acid to be used in various fields such as building, textile, agriculture, food and transportation, since they possess several desirable characteristics such as high latent heat storage capacity (∼158 J/g) and suitable melting points (e.g. in the range 15–17°C). Application of caprylic acid is challenging since it is a liquid at room temperature. Konuklu et al. [70] microencapsulated caprylic acid with various shell materials including UF, MF and urea–melamine–formaldehyde (UMF) by both simple and complex coacervation methods but results demonstrated that caprylic acid could not be microencapsulated via the complex coacervation method. In addition, observation showed that shell of poor quality was produced in the absence of formaldehyde crosslinking agent. They reported that the fabricated microcapsules of this study can be used in TES applications, since they exhibited adequate latent heat storage capacities. Malekipirbazari et al. [71] microencapsulated paraffin wax within a gelatin/acacia shell by the complex coacervation method and reported that the fabricated microcapsules have good prospects in TES applications, since they possessed high latent storage capacity.

3.2.3.3 Sol–gel method

The term sol–gel is the abbreviation of solution-gelation, and it refers to the sol as a stable dispersion of colloidal particles present in the solvent, and gel as the three-dimensional network linked by covalent bonds, and intermolecular forces (e.g. van der Waals forces and hydrogen bonds) composed of the aggregates of the sol particles [14,72]. Sol–gel process is utilised to synthesise various nano-inorganic materials or composite materials by the organic or inorganic materials of metal solution–gelation (“sol–gel”) polymerisation process at low temperatures [73,74] and is considered as a highly flexible method for the synthesis of diverse range of materials, e.g. wide range of nano-/micro-structures [7577], and only requires mild conditions. In this method, a solution (e.g. sol) is initially formed as result of mixing chosen precursor such as a particular metal alkoxide with a solvent, catalyst, and complexing agent. This results in the formation of a stable dispersion of polymers or colloidal particles in chosen solvent as a result of hydrolysis reaction undergone by the precursors. Resulting solution is then added into the PCM oil-in-water emulsion, which results in producing a gel containing discrete microcapsules with PCM droplets entrapped inside them. Inorganic shell materials such as silica and titanium oxide are the most commonly used shell materials. Hydrolysis of precursor compounds such as tetraethoxysilane, sodium silicate or methyl triethoxysilane leads to the formation of the sol solution. During this process, the pH value is maintained very low, usually in the range 2–3, to ensure sol–solution undergoes hydrolysis reaction. Decreasing the pH will promote hydrolysis by protonating the leaving group [75]. Silicate solution gets added into the PCM oil-in-water emulsion drop-by-drop, while continuously stirring the emulsion, to avoid the formation of any continuous gel, instead to encourage the formation of discrete silica gel walls around the PCM droplets. Condensation polymerisation reaction of silica solid particles occurs, and as result, the PCM droplets get surrounded by the silica shells. Figure 7 illustrates the fundamental processing steps involved in the sol–gel method.

Figure 7 
                        Schematic representation of a sol–gel process.
Figure 7

Schematic representation of a sol–gel process.

3.2.4 Chemical encapsulation

3.2.4.1 In situ polymerisation

In this in situ method, droplets are initially formed as a result of dispersing core material (PCM) into an aqueous phase consisting of a small fraction of the emulsifier (Figure 8). This is followed by the addition of adequate monomers or pre-polymers of urea with formaldehyde or melamine with formaldehyde. The pH of the system will then get reduced, which initiates the polycondensation reaction. As result, it yields cross-linked UF or MF resins. Once the resin reaches a high molecular weight, it will become insoluble in the aqueous phase and thus, a precipitate will be formed which is normally observed to deposit at the oil–water interface of the droplet. The resin will eventually harden, and thereby shell of the microcapsule gets formed. Table 2 summarises the properties of microcapsules produced by this method.

Figure 8 
                        Schematic representation of the interfacial polymerisation method employed for the synthesis of microcapsules.
Figure 8

Schematic representation of the interfacial polymerisation method employed for the synthesis of microcapsules.

Table 2

Properties of microcapsules prepared by in situ polymerisation method collected from previous literature

PCM/core material Shell material Properties of microcapsules Ref.
Styrene MF resin
  • Particle size was in the range 20–71 µm

  1. Core material content was ∼60%

[78]
n-Octadecane MF
  • Average diameter of 2.2 µm

  • Average content of 59 wt%

  1. Latent heat of 144 J/g

[79]
n-Octadecane MF
  • Average diameter of 9.2 µm at stirring rate of 4,000 rpm

  • Average content of 70 wt%

  • Latent heat of melting 169 J/g

[80]
n-Octadecane Urea–melamine–formaldehyde
  • Average PCM content of 72 wt%

  • Latent heat of ∼164 J/g

  • Average diameter of 0.8 µm

[81]
n-Nonadecane Urea–melamine–formaldehyde
  • Average PCM content of 69 wt%

  • Latent heat of ∼233 J/g

[82]
n-Eicosane Urea–melamine–formaldehyde
  • Average PCM content of 71 wt%

  • Latent heat of ∼172 J/g

[82]

In general, high reactivity leads to shorter reaction time [83], and this depends greatly upon the type of resin used. Reaction proceeds much faster with MF resins and slower with UF resins. In addition, MF microcapsules have enhanced thermal stability. Park et al. [84] microencapsulated fragrant oil within a UF shell, while Lee et al. [85] microencapsulated floral oil with MF shell material. From latter study, results indicated that the morphology of fabricated microcapsules are influenced mainly by the formaldehyde/melamine molar ratio and the pH of the reaction mixture. Su, Ren and Wang [86] used MF to prepare several types of microcapsules and evaluated their mechanical strengths. Strength of the MF shell reduced as the mass ratio of core/shell increased, and at a core/shell mass ratio of 3:1, the yield point was deduced to be 1.1 × 105 Pa. Salaün et al. [83] microencapsulated n-hexadecane within a MF shell. Results indicated that as the melamine/formaldehyde (MF) molar ratio increased, the surface of the microcapsules became rougher which was attributed by a higher degree of cross-linking of the MF resin, while decreasing the MF molar ratio resulted in producing microcapsules of reduced sizes i.e. a narrow particle distribution was obtained. The poor strength of MF shell can be overcome using amino-aldehyde crosslinking agent, and in addition, its thermal performance will also be improved. The brittleness, and rigidity (e.g. toughness) of this shell can further be enhanced by adding resorcinol as an impact modifier. Zhang and Wang [78] microencapsulated n-octadecane within a resorcinol-modified MF shell. Microcapsules with compact surface was obtained at the core/shell weight ratio of 75/25, and the results showed that these microcapsules exhibited better anti-osmosis property when compared with others. Fang et al. [87] prepared nanocapsules by encapsulating n-tetradecane within an UF shell. They used sodium dodecyl sulphate and resorcin (C6H6O2) as the emulsifier and the system modifier, respectively.

Previous studies had investigated the influence of stirring rate, emulsifier content, cyclohexane on diameter, surface morphology, phase change properties and thermal stabilities of microcapsules produced by this method [80,81,88,89]. Zhang et al. [88] microencapsulated n-octadecane within an UMF copolymer shell. They observed by mixing 30–40% of cyclohexane in the oil phase with heat-treatment under adequate conditions enhanced the thermal stability of synthesised microcapsules significantly by inducing a free space with 5–28% of volume expansion ratio inside the microcapsules [89]. Zhang and Wang [78] reported that microencapsulation efficiency reduced when the amount of styrene–maleic anhydride (SMA) copolymer used increased at a core/shell weight ratio of 80/20 because the shell became too thin, so that it resulted in inducing leakage of n-octadecane. Microcapsules containing 80% of n-octadecane content were successfully produced at a core/shell weight ratio of 75/25, and the optimum microencapsulation efficiency was 92%. Fang et al. [87] reported that using resorcin as crosslinking agent improved the encapsulation efficiency by allowing a high mass content of core material to be encapsulated within the shell. Hence, as the resorcin concentration increased from 0.25 to 5%, the mass ratio of n-tetradecane increased from 30.4 to 61.8%. However, if resorcin concentration was too high, the stickiness of nanocapsules formed increased, and made it more difficult to produce regular nanospheres. Su et al. [90] microencapsulated n-octadecane with methyl methacrylate (MMA) and methacrylic acid (MAA) as shell monomers (e.g. methacrylate-co-methacrylic acid) (PMMA-MAA) and concluded that best microcapsules were produced at shell monomer weight ratio of 80% MMA: 20% MAA. They reported that the microencapsulation efficiency was in the range of 70.4–79.8%, which was due to the high weight percentage of shell that reduced the evaporation of shell monomers significantly. Chen et al. [91] prepared a novel type of octadecylamine-grafted graphene oxide (GO-ODA) – modified microencapsulated PCMs by encapsulating n-octadecane within a MF (MF) shell and reported the encapsulation efficiency to be 88%.

3.2.4.2 Interfacial polymerisation/polycondensation

During this process, the microcapsule wall of the polymer gets formed at the interface present between the two phases provided that each phase contains a suitable reaction monomer. In the beginning, a multifunctional monomer gets dissolved in the organic core material, and the resulting mixture formed in the aqueous phase containing a mixture of emulsifiers and protective colloid stabilisers. Finally, a combination of monomers will be added to the aqueous phase to generate the polymer shell. Figure 9 shows the schematic representation of the processes involved in the interfacial polymerisation method for the encapsulation of PCMs.

Figure 9 
                        Schematic representation of the suspension polymerisation method employed for the synthesis of microcapsules.
Figure 9

Schematic representation of the suspension polymerisation method employed for the synthesis of microcapsules.

In this technique, the wall of the capsule is formed by the means of rapid polymerisation of the hydrophilic and lipophilic monomers, which occurs at the interface of the oil-in-water emulsion. The solution containing a particular lipophilic reactant gets emulsified in the aqueous phase, containing an emulsifier. Then, a particular hydrophilic reactant will be added, in aqueous phase, which results in initiating the interfacial polymerisation, and thereby, a shell will be formed [8]. Once the reaction gets initiated, the wall formed will be turned into a barrier to diffusion, and finally, the rate of interfacial polymerisation reaction gets limited, which results in affecting the morphology and unity of the shell of capsule formed. Interfacial polycondensation, in an emulsion system, is one of the most commonly used chemical methods for the microencapsulation PCMs, due to the advantages such as high reaction speed, mild reaction courses and low permeability. Selection of shell material is very crucial to facilitate the microencapsulated PCMs in various applications because their shell structure determines their chemical and thermal stabilities, release properties, and compatibility. Normally, the structure of the shell material is governed by the monomers, and processing techniques are used. In the past, various shell materials have been investigated, such as melamine formaldehyde (MF) resin, PS, PMMA and UF resin; however, such materials only possess limited applications due to their serious drawback: high chances of emitting toxic gases [92]. In the past, interfacial polymerisation has been used to microencapsulate PCMs with PU shell, due to its formaldehyde-free properties, and it is usually formed by reacting aromatic diisocyanate and aliphatic primary diamines. Unfortunately, this reaction is stated to exhibit serious problems, including production of linear structures, and fast rate of reaction which leads to the generation of microcapsules with poor thermal stability, and compactness [93,94]. Table 3 shows summarises the properties of microcapsules fabricated by this method.

Table 3

Properties of microcapsules prepared by interfacial polymerisation collected from previous literature

PCM/core material Shell material Properties of microcapsules Ref.
Butyl stearate Polyurea
  • Heat of fusion ∼80 J/g

  • Melting point ∼28.7°C

  • Particle diameter ∼20–35 µm

  • Good property of thermal periodicity

  • Good thermal stability after subjection to 400 repetitive heating and cooling cycling

[95]
n-octadecane Silica
  • Particle size 17.0 µm

  • Good thermal stability

  • Enhanced thermal conductivity due to the presence of silica

  • Optimum thermal conductivity (at core/shell weight ratio of 50/50) ∼0.6547 W m−1 K−1

[96]
Octadecane Polyurea
  • Particle size 0.1–1 µm

  • Melting point 28–30°C

  • Latent heat of fusion ∼62.6–112 J/g

  • Heat of crystallization 20.9–22.7°C

[97]
Paraffin RT21 Polyurea (IPDI and EDA as wall monomer)
  • Mean particle diameter was 2.40 µm

  • Latent heat was 92.5 J/g

  • Good anti-osmosis property

  • Good anti-permeability and high stability

[92]
n-Octadecane Polyurea
  • Mean particle size 5–20 µm

  • Jeffamine-based microcapsules exhibited better phase change properties, anti-osmosis property but poorer thermal stabilities

[93]
Butyl stearate Cross-linked network polyurethane with pentaerythritol used as a crosslinking agent
  • Diameter of a microcapsule in the range 10–35 µm

  • Enthalpy was in the range 77–81 J/g. Thus, microcapsules have good latent heat storage performance

  • Enhanced thermal stability due to the presence of network polyurethane shell

[94]

Cho et al. [97] microencapsulated octadecane using toluene-2,4-diisocyanate (TDI), and diethylenetriamine (DETA) as reactive monomers in the disperse phase and aqueous phase, respectively. Multiple reactions resulted in the formation of the PU shell. Polyurea microcapsules had formed as result of the reaction taking place between TDI and DETA, and by the hydrolysis of TDI at the interface, which occurred as result of TDI reacting with amines, in the presence of excess TDI with hydroxyl (OH) group of the non-ionic surfactant (NP-10). Resulting PU microcapsules had an average diameter ranging from 0.1 to 1 µm which is considered as the desirable particle size range for most applications [97]. Su et al. [98] used the same method, and materials along with SMA solid as the dispersant to microencapsulate n-octadecane, and resulting microcapsules exhibited particle size in the range of 1–20 µm. Wei et al. [99] microencapsulated paraffin within a new type of polyamide shell, and prepared microcapsules with an average diameter of 6.4 µm. They found that the resulting microcapsules dispersed in water and organic solvents easily and exhibited good thermal and chemical stability. Zhan et al. [92] encapsulated paraffin within a new type of PU shell which was prepared using isophorone diisocyanate (IPDI) and ethylene diamine (EDA) as shell monomers. Incorporation of IPDI was observed to improve the flexibility of shell formed around the PCM core. Also, the type of emulsifier used influenced the morphology of the microcapsules produced. Hence, microcapsules produced with nonyl phenyl ether (OP-10) as emulsifier exhibited a spherical shape, whereas the ones produced with sodium dodecylbenzenesulfonate as emulsifier had non-spherical structures and adhered to each other. Furthermore, they concluded that the adequate amount of emulsifier required to produce microcapsules of enhanced thermal properties was 1.20 g.

Tamami and Koohmareh [100] prepared new series of PUs derived from 4-aryl-2,6-bis(4-isocyanatophenyl)pyridines by the polycondensation process between 4-aryl-2,6-bis(4-isocyanatophenyl)pyridines with several types of aromatic diamines. They demonstrated successful production of molecule-designed PUs exhibiting good thermal stability; however, the synthesised shell material was observed to possess brittle nature. Zhang and Wang [93] prepared microcapsules by encapsulating n-octadecane within modified-PU shells. They prepared the shells using 2,4-tolyene diisocyanate (TDI) as the oil-soluble monomer, along with three different types of aliphatic amines consisting of various soft segments in the molecular chain, namely EDA, DETA and amine-terminated polyoxypropylene (Jeffamine T403). SEM analysis results showed that microcapsules produced using Jeffamine as the amine monomer exhibited a smoother and compact surface than the ones produced using EDA and DETA. Furthermore, these Jeffamine-based microcapsules were observed to have the largest mean diameter, higher encapsulation efficiency, and better anti-osmosis property. Liang et al. [95] microencapsulated butyl stearate (PCM) by an emulsion system with TDI and EDA monomers and deduced the suitable ratio of weight of core/shell to be 70/30, since encountering higher weight ratios reduced the thermal stability of the microcapsules. Zhan et al. [92] deduced that microcapsules exhibiting a compact structure, good anti-permeability and high stability can be produced at a core/shell ratio of 75/25. Ben Abdelkader et al. [101] microencapsulated neroline fragrance with polyurethane shell using isosorbide and hexane diisocyanate monomers.

Polyurea is considered as one of the most suitable shell materials since polymerisation reaction occurs more rapidly, resulting in producing relatively small and uniform particles. Moreover, PUs are insoluble in water, and they have the potential to incorporate both hydrophilic and hydrophobic core materials. However, they exhibit low stiffness, which is likely to limit the practical applications of the PU microcapsules [102]. Xing et al. [103] endeavoured to enhance the strength of the polymer shell by preparing a series of paraffin double-shell microcapsules exhibiting a relatively low shell permeability. Results showed that reaction between polypropylene glycols and TDI led to the formation of the inner shell, while the outer shell was formed as result of TDI reacting with amines in aqueous phase. Lu et al. [104] used the same method and materials along with SMA solid as the dispersant while butyl stearate was used as the PCM to fabricate microcapsules containing double shell. The stability of these double-shell microcapsules against water-wash and ethanol-wash and thermal resistance was reported to be better than the one of the single-shell microcapsules fabricated by Liang et al. [95]. You et al. [105] observed by adding DVB crosslinking agent, the elastic modulus of shell increased, and the appearance of microcapsules was enhanced when compared with the ones produced in the absence of DVB. Li et al. [106] prepared novel PCM microcapsules with styrene–divinylbenzene copolymer as the inner shell and polyurethane as the outer shell. Both styrene and divinylbenzene was used as co-solvent and shell-forming monomers; however, the microcapsules produced exhibited poor mechanical strength, thus majority of microcapsules obtained was broken.

3.2.4.3 Suspension polymerisation

It is one of the most extensively employed chemical method by researchers for the encapsulation of PCM (see Figure 10) due to its advantages such as its simple, cheap, robust, and environmentally friendly. The basic procedures followed for the microencapsulation of PCMs includes: (1) the polymer monomer gets dissolved into the organic phase (core material), (2) the oil/water emulsion gets formed and (3) the monomer molecules from the core materials get separated and precipitated, and thereby, the solid shell gets formed. This technique is governed by multiple simultaneous mechanisms namely: (1) particle coalescence followed by break-up, (2) secondary nucleation and (3) diffusion of monomer towards the interface. The combined effect of the mentioned mechanisms will determine the size, structure, and surface properties of fabricated microcapsules. In this technique, the degree of encapsulation achieved and the properties of final microcapsules depends mainly upon the complex interaction taking place between the polymer system (e.g. initiator, monomers and polymers) and the continuous phase (e.g. water and suspension agent) present on the surface of the droplets [107]. The morphology of fabricated microcapsules is influenced by the following factors: (1) the mobility present between the components, (2) compatibility between the monomers and polymers, (3) hydrophobicity and (4) the reactivity ratios between the monomers [108]. In general, some suspension polymerisation processes require special precautions to be taken. For example, when using PS, styrene-methyl-poly(divinylbenzene) (PDVB), methacrylate and polyurethane as shell materials nitrogen gas protection facility should be available during the fabrication process, which could take around 5–24 h [90]. Table 4 summarises the final properties of microcapsules resulting from the suspension polymerisation method. Selecting a proper shell material is crucial, in order to microencapsulate PCMs successfully. Intensive work on microencapsulation of PCMs with various shell materials such as organic and inorganic shells are found in the literature. This method only allows the encapsulation of non-polar PCMs. Sánchez et al. [107] demonstrated the possibility of preparing microcapsules with high loading capacity; thus, they successfully fabricated microcapsules with around 50% PCM content by weight (e.g. 50 wt%) but the optimum encapsulation efficiency will change if a different PCM was used. Also, results demonstrated that it is impossible to encapsulate poly(ethylene glycol) within the designated shell due to its hydrophilic nature.

Figure 10 
                        Schematic representation of the emulsion polymerisation method employed for the synthesis of microcapsules.
Figure 10

Schematic representation of the emulsion polymerisation method employed for the synthesis of microcapsules.

Table 4

Outlines the properties of microcapsules obtained by employing a suspension-like polymerisation method

PCM/core material Shell material Properties of microcapsules Ref.
PRS® Paraffin wax PSt
  • Particle diameter ∼30 µm

[109]
Paraffin wax St and MMA
  • Energy storage capacity of 41.65 J/g

  • Latent heat per mass (at MMA/St proportion of 4) ∼84.04 J/g

  • Enhanced thermal conductivity and mechanical strength

[110]
PRS® Paraffin wax St
  • Latent heat ∼41.65 J/g (lower than pure paraffin)

  • Optimum melting heat (nonadecane) was 119.80 J/g

  • Mean diameter of Rubitherm® R20 was the smallest

[107]
Tetradecane
Rubitherm® R27
Rubitherm® R20
Nonadecane
Paraffin MMA and TEVS
  • Enhanced thermal stability, thermal conductivity, and heat transfer rate, i.e. thermal conductivity was reported to reach an increment of 116%

  • Electrical resistance of PCM greater than 1010 Ω/sq

[9]
n-Octadecane St-DVB
  • Thermal decomposition temperature ∼230°C

  • n-Octadecane content (at monomers/n-octadecane mass ratio 1:1) ∼56.8%

  • Optimum enthalpy of 57 J/g (at monomers/n-octadecane mass ratio of 5:2)

  • Average particle size ∼71 µm (at stirring rate of 1,200 rpm)

[105]
Paraffin wax Polystyrene
  • Average particle size was 4.53 μm

  • Average storage capacity (latent heat) was 104.7 J/g

[111]

You et al. [105] microencapsulated n-octadecane (PCM) within a styrene (St)-divinylbenzene (DVB) copolymer shell and observed that as the DVB content increased, the enthalpy of microcapsules decreased. Supatimusro et al. [112] used poly(divinylbenzene) as the shell material instead. Results indicated that shell of microcapsules was dent as the DVB content reduced, thus low DVB content encourages the formation of non-spherical microcapsules. Therefore, they concluded that DVB content influences the shape of the microcapsules produced. Yin et al. [113] deduced that the addition of DVB improved the mechanical strength of the shell; however, incorporating high content of DVB causes serious problems such as cramming of cavities present in the microcapsules with polymeric beads, or difficulty in separating polymer onto the droplet surface during polymerisation. Taguchi et al. [114] microencapsulated n-pentadecane within an MMA shell. Later, Sánchez-Silva et al. [110] microencapsulated paraffin wax within a shell material based on a copolymer of MMA and styrene (St), and results demonstrated the unfeasibility of microencapsulating the PCM when the MMA/St ratio was 2.0, and the monomer/paraffin mass ratio was 3.0. Ma et al. [115] microencapsulated hexadecane (HD) within a shell material resulting from mixing N,N-dimethylaminoethyl methacrylate (DMAEMA) and styrene monomers together. Results showed that the presence of insufficient amount of DMAEMA monomers led to incomplete microencapsulation of HD, and as content of HD increased, it became more difficult for the shell material to encapsulate the PCM. Li et al. [116] microencapsulated n-octadecane with different types of copolymer shells including styrene-1,4-butyleneglycol diacrylate copolymer, styrene-divinylbenzene copolymer, PSDB and PDVB, and the results demonstrated that PSDB shell was unable to prevent penetration of n-octadecane gas through the intact shell when the microcapsule was subjected to heat processing at 200°C for 30 min. Yin et al. [113] used Pickering emulsion polymerisation to microencapsulate dodecanol (PCM) with a new polymer/SiO2 hybrid shell, and the results indicated that the impregnation of SiO2 in the shell led to the enhancement of the mechanical strength of the shell. Thus, the content of SiO2 had to be either 5% or 10% in order to fabricate the microcapsules successfully. In the recent times, the interest of using long-chain acrylate monomers for the microencapsulation of PCMs is growing, mainly for the synthesis of a softer shell. Qiu et al. [117] microencapsulated n-octadecane within a poly(lauryl methacrylate) (PLMA) shell and concluded that these microcapsules had good potential to be used in TES and thermal regulation applications. They found that using pentaerythritol tetraacrylate crosslinking agent enhanced the mechanical strength of the PLMA shell sufficiently. However, increased degree of PLMA shell resulted in reducing its strength, although thermal stability was observed to improve greatly with minor enhancement of the thermal resistance temperature of the microcapsules. Previous literature has reported that the thermal conductivity and TES capacity of organic PCM can be enhanced by adding either nano-materials or minerals. Sahan and Paksoy [118] investigated the role of nanomagnetite (Fe3O4) nanoparticles on the heat transfer properties on PCM (e.g. paraffin) and reported that the addition of this particular type of nanoparticles (particle size in the range of 1–100 nm) [119] is effective in enhancing the heat transfer properties of PCM significantly. Lashgari et al. [120] prepared magnetic microcapsules with n-hexadecane/Fe3O4 as the core material and PMMA as the shell material and reported successful production of microcapsules exhibiting novel dual functional features, such as superparamagnetic properties with low remanence and coercivity.

3.2.4.4 Emulsion polymerisation

Figure 11 shows the basic steps involved in the emulsion polymerisation method. In this method, an insoluble monomer contained in the solvent gets dispersed uniformly by the means of mechanical stirring in the reaction medium which contains a particular emulsifier and surfactant. As result, a polymer membrane gets formed on the surface of the core once the initiator gets added. Thereby, the polymerisation reaction gets initiated in the aqueous phase, which results in the growth of formed polymer particles due to progression of the polymerisation reaction [121]. This method is beneficial for the polymerisation of polymers and can be used to encapsulate designated PCM within a shell; thus, for the synthesis of nanocapsules (e.g. NanoPCM). Alkanes (e.g. paraffins) were the common PCMs used extensively in previous studies, and they were encapsulated within a PS or PMMA shell [122]. Table 5 summarises the properties of microcapsules obtained from this method.

Figure 11 
                        Schematic representation of in situ polymerisation method employed for the synthesis of microcapsules.
Figure 11

Schematic representation of in situ polymerisation method employed for the synthesis of microcapsules.

Table 5

Properties of microcapsules prepared by the emulsion polymerisation method

PCM/core material Shell material Properties of microcapsules Ref.
n-Heptadecane PMMA
  • Microcapsules varied in a narrow range of 0.14–0.40 µm and the average diameter was 0.26 µm (produced at a stirring rate of 2,000 rpm)

  • Good thermal reliability post-exposure to 5,000 repetitive thermal cycling

  • Good thermal stability

  • Latent heats of melting and freezing were 81.5 and −84.7 J/g, respectively, reported to have satisfactory thermal properties for energy storage applications

  • The temperature of melting and freezing was 18.2 and 18.4°C, respectively

  • N-Heptadecane content of ∼38 wt%

[123]
Docosane PMMA
  • Average diameter of 0.16 µm

  • Latent heats of melting and freezing were 54.6 and −48.7 J/g, respectively

  • Encapsulation ratio of docosane was 28 wt%

  • Average latent heats of melting and crystallization were 75.2 and 67.7 J/g, respectively

  • Good chemical stability since no chemical degradation occurred during 5,000 thermal cycling

[124]
n-Octacosane PMMA
  • Average diameter of 0.25 µm at a stirring rate of 2,000 rpm

  • Temperatures of melting and freezing 50.6 and 53.2°C, respectively

  • Latent heats of melting and freezing 86.4 and −88.5 J/g, respectively, and thus, average latent heats of melting and freezing post 5000 thermal cycling was 79.9 and 81.7 J/g, respectively

  • Good chemical stability

  • Average n-octacosane content of 43 wt%

[125]
n-Hexadecane PMMA
  • Mean particle size of 0.22–1.05 µm

[126]
n-Eicosane PMMA
  • Mean diameter of 0.7 μm at a stirring rate of 2,000 rpm

  • Latent heat of melting and freezing was 84.2 and −87.5 J/g

  • 35 wt% content of PCM

[127]
n-Docosane PMMA
  • Average diameter of 0.16 μm

  • Latent heat of 54.6 J/g

  • Encapsulation ratio of 28 wt%

[124]
n-Octadecane MF
  • Encapsulation ratio of ∼40 wt%

  • Average diameter of ∼1.3 μm

  • Latent heat of 166 J/g

  • Microcapsules exhibited a thermal resistant temperature of ∼270°C

[80]

Cho et al. [128] successfully encapsulated n-octadecane within a PS shell to produce nanocapsules of high adhesive property. They used alkali-soluble resin, poly(styrene-co-acrylic acid) as a surfactant and cetyl alcohol as a co-surfactant with DVB as the crosslinking agent. Sari et al. [123] fabricated novel microcapsules by encapsulating n-heptadecane (PCM) within a PMMA (transparent acrylic polymer) (PMMA) shell. Fortuniak et al. [129] used the co-emulsion polymerisation method to microencapsulate n-eicosane within polyhydromethylsiloxane shell. Results demonstrated that microparticles containing n-eicosane core was embedded in the polysiloxane shell. They used 1,3-divinyltetramethyldisiloxane as the crosslinking agent to stabilise the shell and thereby promote the product yield of microcapsules formed. Formaldehyde-based polymers including MF, UF and melamine–urea–formaldehyde are conventionally used for the encapsulation of PCMs in numerous studies, despite possessing a serious drawback; they release formaldehyde residues/fumes that can cause health and environmental problems. Konuklu and Paksoy [130] microencapsulated caprylic (octanoic acid), an eutectic mixture, within a PS shell. Two types of crosslinking agents were used: ethylene glycol dimethacrylate (EGDMA) and allyl methacrylate (AMA) to optimise the strength of the shell formed around the core. PS shell has advantages over the conventionally used shell materials such as they are clear, inexpensive, non-toxic and easy to handle and process. Results demonstrated that type of cross-linking agent used had significant impact on morphology of microcapsules produced. Microcapsules exhibited a uniform geometry was produced with EGDMA, whereas non-spherical or agglomerated microcapsules was observed to be produced with AMA crosslinking agent. Also, increasing the amount of crosslinking agent resulted in decreasing the latent heat capacity of the microcapsules. Overall, both type and quantity of crosslinking agent used influenced the microencapsulation efficiency greatly. Numerous studies on the encapsulation of traditional organic PCMs such as paraffins and paraffin waxes are available since it is challenging to encapsulate fatty acids (e.g. eutectic mixtures) because they contain reactive functional groups; however, a limited number of studies have demonstrated the possibility of encapsulating fatty acids via the emulsion polymerisation method. Sari [131] prepared micro-nanocapsules successfully by encapsulating three different types of fatty acids such as capric acid, lauric acid and myristic acid within a PS shell via the emulsion polymerisation method. The encapsulation ratios obtained were 45.4, 43.56 and 48.7 wt%, respectively. In a later study, Sari, Alkan and Özcan [132] prepared nano-/micro-capsules by encapsulating capric-stearic acid eutectic mixture (C-SEM) within a PMMA shell.

4 Conclusions

PCMs have number of benefits; however, the leakage problem limits their application greatly. This shortcoming can be overcome through the employment of the nano-/micro-encapsulation technologies. Current study has extensively reviewed the specifications of PCMs, encapsulation technology/methods, and provided a summary of properties of nano-/micro-capsules fabricated in previous studies. The encapsulation methods of PCM used to overcome the leakage problem of solid–liquid PCMs have been discussed to get a better understanding of factors likely to influence the final properties of synthesised nano-/micro-capsules. Nano-/micro-encapsulation technologies have gained considerable attention, mainly because it allows for more flexible choices of shell materials and PCM selection according to the practical requirements. Major shortcomings of this technology include high processing costs, and selecting the correct materials. Therefore, to widen the potential usage of this technology, it is necessary to reduce the overall encapsulation costs, facilitate production methods, and enhance the stabilisation of PCMs. Most studies have focused on microencapsulating PCM within synthetic petroleum-based polymers; however, it will be more beneficial to develop and apply nanoencapsulation techniques for the synthesis of nanocapsules, since they have number of desirable characteristics, including better loading capacity and encapsulation efficiency, improved structural stability and fracture resistances, when compared with microcapsules resulting from microencapsulation. To promote the sustainability of the encapsulation technology, it is necessary to utilise biopolymers such as cellulose/nanocellulose as shell materials because such polymers are environmentally friendly and are available abundantly.

  1. Conflict of interest: The authors declare no conflict of interest regarding the publication of this paper.

References

[1] Huo X, Li W, Wang Y, Han N, Wang J, Wang N, et al. Chitosan composite microencapsulated comb-like polymeric phase change material via coacervation microencapsulation. Carbohydr Polym. 2018;200:602–10.10.1016/j.carbpol.2018.08.003Search in Google Scholar PubMed

[2] Souayfane F, Fardoun F, Biwole PH. Phase change materials (PCM) for cooling applications in buildings: a review. Energy Build. 2016;129:396–431.10.1016/j.enbuild.2016.04.006Search in Google Scholar

[3] Mondal S. Phase change materials for smart textiles – an overview. Appl Therm Eng. 2008;28:1536–50.10.1016/j.applthermaleng.2007.08.009Search in Google Scholar

[4] Kee SY, Munusamy Y, Ong KS. Review of solar water heaters incorporating solid–liquid organic phase change materials as thermal storage. Appl Therm Eng. 2018;131:455–71.10.1016/j.applthermaleng.2017.12.032Search in Google Scholar

[5] Yang J, Qi GQ, Liu Y, Bao RY, Liu ZY, Yang W, et al. Hybrid graphene aerogels/phase change material composites: thermal conductivity, shape-stabilization and light-to-thermal energy storage. Carbon. 2016;100:693–702.10.1016/j.carbon.2016.01.063Search in Google Scholar

[6] Abdelsalam MY, Teamah HM, Lightstone MF, Cotton JS. Hybrid thermal energy storage with phase change materials for solar domestic hot water applications: direct versus indirect heat exchange systems. Renewable Energy. 2020;147:77–88.10.1016/j.renene.2019.08.121Search in Google Scholar

[7] Gunasekara SN, Martin V, Chiu JN. Phase equilibrium in the design of phase change materials for thermal energy storage: state-of-the-art. Renewable Sustainable Energy Rev. 2017;73:558–81.10.1016/j.rser.2017.01.108Search in Google Scholar

[8] Jamekhorshid A, Sadrameli SM, Farid M. A review of microencapsulation methods of phase change materials (PCMs) as a thermal energy storage (TES) medium. Renewable Sustainable Energy Rev. 2014;31:531–42.10.1016/j.rser.2013.12.033Search in Google Scholar

[9] Ng DQ, Tseng YL, Shih YF, Lian HY, Yu YH. Synthesis of novel phase change material microcapsule and its application. Polymer. 2017;133:250–62.10.1016/j.polymer.2017.11.046Search in Google Scholar

[10] Umair MM, Zhang Y, Iqbal K, Zhang S, Tang B. Novel strategies and supporting materials applied to shape-stabilize organic phase change materials for thermal energy storage – a review. ppl Energy. 2019;235:846–73.10.1016/j.apenergy.2018.11.017Search in Google Scholar

[11] Amaral C, Vicente R, Marques PAAP, Barros-Timmons A. Phase change materials and carbon nanostructures for thermal energy storage: a literature review. Renewable Sustainable Energy Rev. 2017;79:1212–28.10.1016/j.rser.2017.05.093Search in Google Scholar

[12] Kahraman D, Biçer A, Sar A, Alkan C. Microencapsulated n-alkane eutectics in polystyrene for solar thermal applications. Sol Energy. 2018;160:32–42.10.1016/j.solener.2017.11.072Search in Google Scholar

[13] Eanest Jebasingh B, Valan Arasu A. A comprehensive review on latent heat and thermal conductivity of nanoparticle dispersed phase change material for low-temperature applications. Energy Storage Mater. 2020;24:52–74.10.1016/j.ensm.2019.07.031Search in Google Scholar

[14] Milián YE, Gutiérrez A, Grágeda M, Ushak S. A review on encapsulation techniques for inorganic phase change materials and the in fluence on their thermophysical properties. Renewable Sustainable Energy Rev. 2017;73:983–99.10.1016/j.rser.2017.01.159Search in Google Scholar

[15] Sharma A, Tyagi VV, Chen CR, Buddhi D. Review on thermal energy storage with phase change materials and applications. Renewable Sustainable Energy Rev. 2009;13:318–45.10.1016/j.rser.2007.10.005Search in Google Scholar

[16] Pomianowski M, Heiselberg P, Zhang Y. Review of thermal energy storage technologies based on PCM application in buildings. Energy Build. 2013;67:56–69.10.1016/j.enbuild.2013.08.006Search in Google Scholar

[17] Saafi K, Daouas N. Energy and cost efficiency of phase change materials integrated in building envelopes under Tunisia Mediterranean climate. Energy. 2019;187:1159–87.10.1016/j.energy.2019.115987Search in Google Scholar

[18] Fabiani C, Piselli C, Pisello AL. Thermo-optic durability of cool roof membranes: effect of shape stabilized phase change material inclusion on building energy efficiency. Energy Build. 2020;207:109592.10.1016/j.enbuild.2019.109592Search in Google Scholar

[19] Kasaeian A, Pourfayaz F, Khodabandeh E. Experimental studies on the applications of PCMs and nano-PCMs in buildings: a critical review. Energy Build. 2017;154:96–112.10.1016/j.enbuild.2017.08.037Search in Google Scholar

[20] Deng Y, Yang L. Preparation and characterization of polyethylene glycol (PEG) hydrogel as shape-stabilized phase change material. Appl Therm Eng. 2017;114:1014–7.10.1016/j.applthermaleng.2016.11.207Search in Google Scholar

[21] Berardi U, Gallardo AA. Properties of concretes enhanced with phase change materials for building applications. Energy Build. 2019;199:402–14.10.1016/j.enbuild.2019.07.014Search in Google Scholar

[22] Liu L, Peng B, Yue C, Guo M, Zang M. Low-cost, shape-stabilized fly ash composite phase change material synthesized by using a facile process for building energy efficiency. Mater Chem Phys. 2019;222:87–95.10.1016/j.matchemphys.2018.09.072Search in Google Scholar

[23] Park JH, Kang Y, Lee J, Wi S, Chang JD, Kim S. Analysis of walls of functional gypsum board added with porous material and phase change material to improve hygrothermal performance. Energy Build. 2019;183:803–16.10.1016/j.enbuild.2018.11.023Search in Google Scholar

[24] Liu L, Xiong B, Men Y. Facile preparation of porous plaster board containing phase change capsules using gel template. Energy Build. 2017;156:134–9.10.1016/j.enbuild.2017.09.074Search in Google Scholar

[25] Pasupathy A, Velraj R, Seeniraj RV. Phase change material-based building architecture for thermal management in residential and commercial establishments. Renewable Sustainable Energy Rev. 2008;12:39–64.10.1016/j.rser.2006.05.010Search in Google Scholar

[26] Zhang Y, Zheng S, Zhu S, Ma J, Sun Z, Farid M. Evaluation of paraffin infiltrated in various porous silica matrices as shape-stabilized phase change materials for thermal energy storage. Energy Convers Manag. 2018;171:361–70.10.1016/j.enconman.2018.06.002Search in Google Scholar

[27] Zahir MH, Mohamed SA, Saidur R, Al-Sulaiman FA. Supercooling of phase-change materials and the techniques used to mitigate the phenomenon. Appl Energy. 2019;240:793–817.10.1016/j.apenergy.2019.02.045Search in Google Scholar

[28] Liu Z, Yu Z, (Jerry), Yang T, Qin D, Li S, Zhang G, et al. A review on macro-encapsulated phase change material for building envelope applications. Build Environ. 2018;144:281–94.10.1016/j.buildenv.2018.08.030Search in Google Scholar

[29] Su W, Darkwa J, Kokogiannakis G. Review of solid–liquid phase change materials and their encapsulation technologies. Renewable Sustainable Energy Rev. 2015;48:373–91.10.1016/j.rser.2015.04.044Search in Google Scholar

[30] Akeiber H, Nejat P, Majid MZA, Wahid MA, Jomehzadeh F, Zeynali Famileh I, et al. A review on phase change material (PCM) for sustainable passive cooling in building envelopes. Renewable Sustainable Energy Rev. 2016;60:1470–97.10.1016/j.rser.2016.03.036Search in Google Scholar

[31] Wahid MA, Hosseini SE, Hussen HM, Akeiber HJ, Saud SN. An overview of phase change materials for construction architecture thermal management in hot and dry climate region. Appl Therm Eng. 2017;112:1240–59.10.1016/j.applthermaleng.2016.07.032Search in Google Scholar

[32] Fallahi A, Guldentops G, Tao M, Granados-Focil S, Van Dessel S. Review on solid–solid phase change materials for thermal energy storage: Molecular structure and thermal properties. Appl Therm Eng. 2017;127:1427–41.10.1016/j.applthermaleng.2017.08.161Search in Google Scholar

[33] Wilhelm E, Richter C, Rapp BE. Phase change materials in microactuators: basics, applications and perspectives. Sens Actuators A. 2018;271:303–47.10.1016/j.sna.2018.01.043Search in Google Scholar

[34] Khan Z, Khan Z, Ghafoor A. A review of performance enhancement of PCM based latent heat storage system within the context of materials, thermal stability and compatibility. Energy Convers Manag. 2016;115:132–58.10.1016/j.enconman.2016.02.045Search in Google Scholar

[35] Salunkhe PB, Shembekar PS. A review on effect of phase change material encapsulation on the thermal performance of a system. Renewable Sustainable Energy Rev. 2012;16:5603–16.10.1016/j.rser.2012.05.037Search in Google Scholar

[36] Mishra A, Shukla A, Sharma A. Latent heat storage through phase change materials. Resonance. 2015;20:532–41.10.1007/s12045-015-0212-5Search in Google Scholar

[37] Bayon A, Liu M, Sergeev D, Grigore M, Bruno F, Muller M. Novel solid–solid phase-change cascade systems for high-temperature thermal energy storage. Sol Energy. 2019;177:274–83.10.1016/j.solener.2018.10.085Search in Google Scholar

[38] Silva T, Vicente R, Rodrigues F. Literature review on the use of phase change materials in glazing and shading solutions. Renewable Sustainable Energy Rev. 2016;53:515–35.10.1016/j.rser.2015.07.201Search in Google Scholar

[39] Huang X, Guo J, He J, Gong Y, Wang D, Song Z. Novel phase change materials based on fatty acid eutectics and triallyl isocyanurate composites for thermal energy storage. J Appl Polym Sci. 2017;134:1–8.10.1002/app.44866Search in Google Scholar

[40] Madene A, Jacquot M. Review flavour encapsulation and controlled release – a review. 2006;41:1–21.10.1111/j.1365-2621.2005.00980.xSearch in Google Scholar

[41] Hawlader MNA, Uddin MS, Khin MM. Microencapsulated PCM thermal-energy storage system. Appl Energy. 2003;74:195–202.10.1016/S0306-2619(02)00146-0Search in Google Scholar

[42] Comunian TA, Favaro-Trindade CS. Microencapsulation using biopolymers as an alternative to produce food enhanced with phytosterols and omega-3 fatty acids: a review. Food Hydrocolloids. 2016;61:442–57.10.1016/j.foodhyd.2016.06.003Search in Google Scholar

[43] Arán-Ais F, Pérez-Limiñana MÁ, Sánchez-Navarro MM, Orgilés-Barceló C. Developments in microencapsulation technology to improve adhesive formulations. J Adhes. 2012;88:391–405.10.1080/00218464.2012.660368Search in Google Scholar

[44] Magdalena A, Pop A, Cimpeanu C, Turcus V. European journal of medicinal chemistry nanoencapsulation techniques for compounds and products with antioxidant and antimicrobial activity – a critical view. Eur J Med Chem. 2018;157:1326–45.10.1016/j.ejmech.2018.08.076Search in Google Scholar PubMed

[45] Chen W, Lv G, Hu W, Li D, Chen S, Dai Z. Synthesis and applications of graphene quantum dots: a review. Nanotechnol Rev. 2018;7:157–85.10.1515/ntrev-2017-0199Search in Google Scholar

[46] Zangabad PS, Mirkiani S, Shahsavari S, Masoudi B, et al. Stimulus-responsive liposomes as smart nanoplatforms for drug delivery applications. Nanotechnol Rev. 2018;7:95–122.10.1515/ntrev-2017-0154Search in Google Scholar PubMed PubMed Central

[47] Li J, Yao M, Shao Y, Yao D. The application of bio-nanotechnology in tumor diagnosis and treatment: a view. Nanotechnol Rev. 2018;7:257–66.10.1515/ntrev-2018-0011Search in Google Scholar

[48] Martins IM, Barreiro MF, Coelho M, Rodrigues AE. Microencapsulation of essential oils with biodegradable polymeric carriers for cosmetic applications. Chem Eng J. 2014;245:191–200.10.1016/j.cej.2014.02.024Search in Google Scholar

[49] Tyagi VV, Kaushik SC, Tyagi SK, Akiyama T. Development of phase change materials based microencapsulated technology for buildings: a review. Renewable Sustainable Energy Rev. 2011;15:1373–91.10.1016/j.rser.2010.10.006Search in Google Scholar

[50] Sawalha H, Schroën K, Boom R. Biodegradable polymeric microcapsules: preparation and properties. Chem Eng J. 2011;169:1–10.10.1016/j.cej.2011.02.078Search in Google Scholar

[51] Ahangaran F, Navarchian AH, Picchioni F. Material encapsulation in poly(methyl methacrylate) shell: a review. Appl Polym Sci. 2019;48039:1–21.10.1002/app.48039Search in Google Scholar

[52] Kazemi M, Ghobadi M, Mirzaie A. Cobalt ferrite nanoparticles (CoFe2O4 MNPs) as catalyst and support: magnetically recoverable nanocatalysts in organic synthesis. Nanotechnol Rev. 2018;7:43–68.10.1515/ntrev-2017-0138Search in Google Scholar

[53] Alva G, Lin Y, Liu L, Fang G. Synthesis, characterization and applications of microencapsulated phase change materials in thermal energy storage: a review. Energy Build. 2017;144:276–94.10.1016/j.enbuild.2017.03.063Search in Google Scholar

[54] Samantha SC, Bruna ASM, Adriana RM, Fabio B, Sandro AR, Aline RCA. Drying by spray drying in the food industry: micro-encapsulation, process parameters and main carriers used. Afr J Food Sci. 2015;9:462–70.10.5897/AJFS2015.1279Search in Google Scholar

[55] Arpagaus C, Collenberg A, Rütti D, Assadpour E, Jafari SM. Nano spray drying for encapsulation of pharmaceuticals. Int J Pharm. 2018;546:194–214.10.1016/j.ijpharm.2018.05.037Search in Google Scholar PubMed

[56] Rajabi H, Ghorbani M, Jafari SM, Sadeghi Mahoonak A, Rajabzadeh G. Retention of saffron bioactive components by spray drying encapsulation using maltodextrin, gum Arabic and gelatin as wall materials. Food Hydrocolloids. 2015;51:327–37.10.1016/j.foodhyd.2015.05.033Search in Google Scholar

[57] Shu B, Yu W, Zhao Y, Liu X. Study on microencapsulation of lycopene by spray-drying. J Food Eng. 2006;76:664–9.10.1016/j.jfoodeng.2005.05.062Search in Google Scholar

[58] Ziaee A, Albadarin AB, Padrela L, Femmer T, O’Reilly E, Walker G. Spray drying of pharmaceuticals and biopharmaceuticals: critical parameters and experimental process optimization approaches. Eur J Pharm Sci. 2019;127:300–18.10.1016/j.ejps.2018.10.026Search in Google Scholar PubMed

[59] Borreguero AM, Valverde JL, Rodríguez JF, Barber AH, Cubillo JJ, Carmona M. Synthesis and characterization of microcapsules containing Rubitherm®RT27 obtained by spray drying. Chem Eng J. 2011;166:384–90.10.1016/j.cej.2010.10.055Search in Google Scholar

[60] Fei B, Lu H, Qi K, Shi H, Liu T, Li X, et al. Multi-functional microcapsules produced by aerosol reaction. J Aerosol Sci. 2008;39:1089–98.10.1016/j.jaerosci.2008.07.007Search in Google Scholar

[61] Rai M, Pandit R, Gaikwad S, Yadav A, Gade A. Potential applications of curcumin and curcumin nanoparticles: from traditional therapeutics to modern nanomedicine. Nanotechnol Rev. 2015;4:161–72.10.1515/ntrev-2015-0001Search in Google Scholar

[62] Ozkan G, Franco P, De Marco I, Xiao J, Capanoglu E. A review of microencapsulation methods for food antioxidants: principles, advantages, drawbacks and applications. Food Chem. 2019;272:494–506.10.1016/j.foodchem.2018.07.205Search in Google Scholar PubMed

[63] Timilsena YP, Akanbi TO, Khalid N, Adhikari B, Barrow CJ. Complex coacervation: principles, mechanisms and applications in microencapsulation. Int J Biol Macromol. 2019;121:1276–86.10.1016/j.ijbiomac.2018.10.144Search in Google Scholar PubMed

[64] Yasmin R, Shah M, Khan SA, Ali R. Gelatin nanoparticles: a potential candidate for medical applications. Nanotechnol Rev. 2017;6:191–207.10.1515/ntrev-2016-0009Search in Google Scholar

[65] Hawlader MNA, Uddin MS, Zhu HJ. Preparation and evaluation of a novel solar storage material: microencapsulated paraffin. Int J Sol Energy. 2000;20:227–38.10.1080/01425910008914357Search in Google Scholar

[66] Onder E, Sarier N, Cimen E. Encapsulation of phase change materials by complex coacervation to improve thermal performances of woven fabrics. Thermochim Acta. 2007;467:63–72.10.1016/j.tca.2007.11.007Search in Google Scholar

[67] Deveci SS, Basal G. Preparation of PCM microcapsules by complex coacervation of silk fibroin and chitosan. Colloid Polym Sci. 2009;287:1455–67.10.1007/s00396-009-2115-zSearch in Google Scholar

[68] Jun-xia X, Hai-yan Y, Jian Y. Microencapsulation of sweet orange oil by complex coacervation with soybean protein isolate/gum Arabic. Food Chem. 2011;125:1267–72.10.1016/j.foodchem.2010.10.063Search in Google Scholar

[69] Bayés-García L, Ventolà L, Cordobilla R, Benages R, Calvet T, Cuevas-Diarte MA. Phase Change Materials (PCM) microcapsules with different shell compositions: Preparation, characterization and thermal stability. Sol Energy Mater Sol Cell. 2010;94:1235–40.10.1016/j.solmat.2010.03.014Search in Google Scholar

[70] Konuklu Y, Unal M, Paksoy HO. Microencapsulation of caprylic acid with different wall materials as phase change material for thermal energy storage. Sol Energy Mater Sol Cell. 2014;120:536–42.10.1016/j.solmat.2013.09.035Search in Google Scholar

[71] Malekipirbazari M, Sadrameli SM, Dorkoosh F, Sharifi H. Synthetic and physical characterization of phase change materials microencapsulated by complex coacervation for thermal energy storage applications. Int J Energy Res. 2014;31:135–47.10.1002/er.3153Search in Google Scholar

[72] Ahmadzadeh MA, Chini SF, Sadeghi A. Size and shape tailored sol–gel synthesis and characterization of lanthanum phosphate (LaPO4) nanoparticles. Mater Des. 2019;181:108058.10.1016/j.matdes.2019.108058Search in Google Scholar

[73] Jia F, Li G, Yang B, Yu B, Shen Y, Cong H. Investigation of rare earth upconversion fluorescent nanoparticles in biomedical field. Nanotechnol Rev. 2019;8:1–17.10.1515/ntrev-2019-0001Search in Google Scholar

[74] Ali A, Phull AR, Zia M. Elemental zinc to zinc nanoparticles: is ZnO NPs crucial for life? Synthesis, toxicological, and environmental concerns. Nanotechnol Rev. 2018;7:413–41.10.1515/ntrev-2018-0067Search in Google Scholar

[75] Owens GJ, Singh RK, Foroutan F, Alqaysi M, Han C, Mahapatra C, et al. Sol–gel based materials for biomedical applications. Prog Mater Sci. 2016;77:1–79.10.1016/j.pmatsci.2015.12.001Search in Google Scholar

[76] Ban I, Stergar J, Maver U. NiCu magnetic nanoparticles: review of synthesis methods, surface functionalization approaches, and biomedical applications. Nanotechnol Rev. 2018;7:187–207.10.1515/ntrev-2017-0193Search in Google Scholar

[77] Das S, Srivastava VC. An overview of the synthesis of CuO–ZnO nanocomposite for environmental and other applications. Nanotechnol Rev. 2018;7:267–82.10.1515/ntrev-2017-0144Search in Google Scholar

[78] Zhang H, Wang X. Fabrication and performances of microencapsulated phase change materials based on n-octadecane core and resorcinol-modified melamine-formaldehyde shell. Colloids Surf A. 2009;332:129–38.10.1016/j.colsurfa.2008.09.013Search in Google Scholar

[79] Li W, Zhang XX, Wang XC, Niu JJ. Preparation and characterization of microencapsulated phase change material with low remnant formaldehyde content. Mater Chem Phys. 2007;106:437–42.10.1016/j.matchemphys.2007.06.030Search in Google Scholar

[80] Zhang XX, Fan YF, Tao XM, Yick KL. Fabrication and properties of microcapsules and nanocapsules containing n-octadecane. Mater Chem Phys. 2004;88:300–7.10.1016/j.matchemphys.2004.06.043Search in Google Scholar

[81] Zhang XX, Tao XM, Yick KL, Fan YF. Expansion space and thermal stability of microencapsulated n-octadecane. J Appl Polym Sci. 2005;97:390–6.10.1002/app.21760Search in Google Scholar

[82] Zhang XX, Fan YF, Tao XM, Yick KL. Crystallization and prevention of supercooling of microencapsulated n-alkanes. J Colloid Interface Sci. 2005;281:299–306.10.1016/j.jcis.2004.08.046Search in Google Scholar PubMed

[83] Salaün F, Devaux E, Bourbigot S, Rumeau P. Influence of process parameters on microcapsules loaded with n-hexadecane prepared by in situ polymerization. Chem Eng J. 2009;155:457–65.10.1016/j.cej.2009.07.018Search in Google Scholar

[84] Park S-J, Shin Y-S, Lee J-R. Preparation and characterization of microcapsules containing lemon oil. J Colloid Interface Sci. 2001;6:1243–51.10.1006/jcis.2001.7727Search in Google Scholar

[85] Lee HY, Lee SJ, Cheong IW, Kim JH. Microencapsulation of fragrant oil via in situ polymerization: effects of pH and melamine-formaldehyde molar ratio. J Microencapsul. 2002;19:559–69.10.1080/02652040210140472Search in Google Scholar PubMed

[86] Su J, Ren L, Wang L. Preparation and mechanical properties of thermal energy storage microcapsules. Colloid Polym Sci. 2005;284:224–8.10.1007/s00396-005-1368-4Search in Google Scholar

[87] Fang G, Li H, Yang F, Liu X, Wu S. Preparation and characterization of nano-encapsulated n-tetradecane as phase change material for thermal energy storage. Chem Eng J. 2009;153:217–21.10.1016/j.cej.2009.06.019Search in Google Scholar

[88] Zhang XX, Tao XM, Yick KL, Wang XC. Structure and thermal stability of microencapsulated phase-change materials. Colloid Polym Sci. 2004;282:330–6.10.1007/s00396-003-0925-ySearch in Google Scholar

[89] Fan YF, Zhang XX, Wu SZ, Wang XC. Thermal stability and permeability of microencapsulated n-octadecane and cyclohexane. Thermochim Acta. 2005;429:25–29.10.1016/j.tca.2004.11.025Search in Google Scholar

[90] Su W, Darkwa J, Kokogiannakis G, Zhou T, Li Y. Preparation of microencapsulated phase change materials (MEPCM) for thermal energy storage. Energy Procedia. 2017;121:95–101.10.1016/j.egypro.2017.07.485Search in Google Scholar

[91] Chen DZ, Qin SY, Tsui GCP, Tang CY, Ouyang X, Liu JH, et al. Fabrication, morphology and thermal properties of octadecylamine-grafted graphene oxide-modified phase-change microcapsules for thermal energy storage. Composites Part B. 2019;157:239–47.10.1016/j.compositesb.2018.08.066Search in Google Scholar

[92] Zhan S, Chen S, Chen L, Hou W. Preparation and characterization of polyurea microencapsulated phase change material by interfacial polycondensation method. Powder Technol. 2016;292:217–22.10.1016/j.powtec.2016.02.007Search in Google Scholar

[93] Zhang H, Wang X. Synthesis and properties of microencapsulated n-octadecane with polyurea shells containing different soft segments for heat energy storage and thermal regulation. Sol Energy Mater Sol Cell. 2009;93:1366–76.10.1016/j.solmat.2009.02.021Search in Google Scholar

[94] Lu S, Shen T, Xing J, Song Q, Shao J, Zhang J, et al. Preparation and characterization of cross-linked polyurethane shell microencapsulated phase change materials by interfacial polymerization. Mater Lett. 2018;211:36–39.10.1016/j.matlet.2017.09.074Search in Google Scholar

[95] 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;50:723–9.10.1016/j.enconman.2008.09.044Search in Google Scholar

[96] Zhang H, Sun S, Wang X, Wu D. Fabrication of microencapsulated phase change materials based on n-octadecane core and silica shell through interfacial polycondensation. Colloids Surf A. 2011;389:104–17.10.1016/j.colsurfa.2011.08.043Search in Google Scholar

[97] 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;280:260–6.10.1007/s00396-001-0603-xSearch in Google Scholar

[98] Su JF, Wang LX, Ren L. Synthesis of polyurethane microPCMs containing n-octadecane by interfacial polycondensation: influence of styrene-maleic anhydride as a surfactant. Colloids Surf A. 2007;299:268–75.10.1016/j.colsurfa.2006.11.051Search in Google Scholar

[99] Wei J, Li Z, Liu L, Liu X. Preparation and characterization of novel polyamide paraffin MEPCM by interfacial polymerization technique. J Appl Polym Sci. 2013;127:4588–93.10.1002/app.37681Search in Google Scholar

[100] Tamami B, Koohmareh GA. Synthesis and characterization of polyureas derived from 4-aryl-2,6-bis(4-isocyanatophenyl)pyridines. Des Monomers Polym. 2007;10:221–33.10.1163/156855507780949236Search in Google Scholar

[101] Ben Abdelkader M, Azizi N, Baffoun A, Chevalier Y, Majdoub M. New microcapsules based on isosorbide for cosmetotextile: preparation and characterization. Ind Crop Prod. 2018;123:591–9.10.1016/j.indcrop.2018.07.020Search in Google Scholar

[102] Liu F, Wang J, Qian X. Integrating phase change materials into concrete through microencapsulation using cenospheres. Cem Concr Compos. 2017;80:317–25.10.1016/j.cemconcomp.2017.04.001Search in Google Scholar

[103] Xing J, Li Y, Newton E, Yeung K-W. Method for encapsulating phase transitional paraffin compound that can undergo phase transition and microcapsule resulting therefrom. 2006, United States Patents, p. 099427.Search in Google Scholar

[104] Lu S, Xing J, Zhang Z, Jia G. Preparation and characterization of polyurea/polyurethane double-shell microcapsules containing butyl stearate through interfacial polymerization. J Appl Polym Sci. 2011;116:2658–67.10.1002/app.33994Search in Google Scholar

[105] You M, Wang X, Zhang X, Zhang L, Wang J. Microencapsulated n-octadecane with styrene-divinybenzene co-polymer shells. J Polym Res. 2011;18:49–58.10.1007/s10965-010-9390-8Search in Google Scholar

[106] Li W, Zhang XX, Wang XC, Tang GY, Shi HF. Fabrication and morphological characterization of microencapsulated phase change materials (MicroPCMs) and macrocapsules containing MicroPCMs for thermal energy storage. Energy. 2012;38:249–54.10.1016/j.energy.2011.12.005Search in Google Scholar

[107] Sánchez L, Sánchez P, de Lucas A, Carmona M, Rodriguez JF. Microencapsulation of PCMs with a polystyrene shell. Colloid Polym Sci. 2007;285:1377–85.10.1007/s00396-007-1696-7Search in Google Scholar

[108] Stubbs JM, Sundberg DC. Core-shell and other multiphase latex particles-confirming their morphologies and relating those to synthesis variables. J Coat Technol Res. 2008;5:169–80.10.1007/s11998-007-9060-xSearch in Google Scholar

[109] Sánchez L, Sánchez P, Carmona M, de Lucas A, Rodriguez JF. Influence of operation conditions on the microencapsulation of PCMs by means of suspension-like polymerization. Colloid Polym Sci. 2008;286:1019–27.10.1007/s00396-008-1864-4Search in Google Scholar

[110] Sánchez-Silva L, Rodríguez JF, Romero A, Borreguero AM, Carmona M, Sanchez P. Microencapsulation of PCMs with a styrene-methyl methacrylate copolymer shell by suspension-like polymerisation. Chem Eng J. 2010;157:216–22.10.1016/j.cej.2009.12.013Search in Google Scholar

[111] Sánchez P, Sánchez-Fernandez MV, Romero A, Rodríguez JF, Sanchez-Silva L. Development of thermo-regulating textiles using paraffin wax microcapsules. Thermochim Acta. 2010;498:16–21.10.1016/j.tca.2009.09.005Search in Google Scholar

[112] Supatimusro D, Promdsorn S, Thipsit S, Boontung W, Chaiyasat P, Chaiyasat A. Poly(divinylbenzene) microencapsulated octadecane for use as a heat storage material: influences of microcapsule size and monomer/octadecane ratio. Polym - Plast Technol Eng. 2012;51:1167–72.10.1080/03602559.2012.694949Search in Google Scholar

[113] Yin D, Ma L, Liu J, Zhang Q. Pickering emulsion: a novel template for microencapsulated phase change materials with polymer-silica hybrid shell. Energy. 2014;64:575–81.10.1016/j.energy.2013.10.004Search in Google Scholar

[114] Taguchi Y, Yokoyama H, Kado H, Tanaka M. Preparation of PCM microcapsules by using oil absorbable polymer particles. Colloids Surf A. 2007;301:41–7.10.1016/j.colsurfa.2006.12.019Search in Google Scholar

[115] Ma GH, Su ZG, Omi S, Sundberg D, Stubbs J. Microencapsulation of oil with poly(styrene-N,N-dimethylaminoethyl methacrylate) by SPG emulsification technique: Effects of conversion and composition of oil phase. J Colloid Interface Sci. 2003;266:282–94.10.1016/S0021-9797(03)00692-1Search in Google Scholar

[116] Li W, Song G, Tang G, Chu X, Ma S, Liu C. Morphology, structure and thermal stability of microencapsulated phase change material with copolymer shell. Energy. 2011;36:785–91.10.1016/j.energy.2010.12.041Search in Google Scholar

[117] Qiu X, Lu L, Wang J, Tang G, Song G. Fabrication, thermal properties and thermal stabilities of microencapsulated n-alkane with poly(lauryl methacrylate) as shell. Thermochim Acta. 2015;620:10–17.10.1016/j.tca.2015.10.001Search in Google Scholar

[118] Sahan N, Paksoy HO. Thermal enhancement of paraffin as a phase change material with nanomagnetite. Sol Energy Mater Sol Cell. 2014;126:56–61.10.1016/j.solmat.2014.03.018Search in Google Scholar

[119] Scott A, Vadalasetty KP, Chwalibog A, Sawosz E. Copper nanoparticles as an alternative feed additive in poultry diet: a review. Nanotechnol Rev. 2018;7:69–93.10.1515/ntrev-2017-0159Search in Google Scholar

[120] Lashgari S, Mahdavian AR, Arabi H, Ambrogi V, Marturano V. Preparation of acrylic PCM microcapsules with dual responsivity to temperature and magnetic field changes. Eur Polym J. 2018;101:18–28.10.1016/j.eurpolymj.2018.02.011Search in Google Scholar

[121] Koehler JM, Visaveliya N, Knauer A. Controlling formation and assembling of nanoparticles by control of electrical charging, polarization, and electrochemical potential. Nanotechnol Rev. 2014;3:553–68.10.1515/ntrev-2014-0006Search in Google Scholar

[122] 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;13:814–26.10.1016/j.nanoen.2015.02.016Search in Google Scholar

[123] Sari A, Alkan C, Karaipekli A. Preparation, characterization and thermal properties of PMMA/n-heptadecane microcapsules as novel solid–liquid microPCM for thermal energy storage. Appl Energy. 2010;87:1529–34.10.1016/j.apenergy.2009.10.011Search in Google Scholar

[124] Alkan C, Sari A, Karaipekli A, Uzun O. Preparation, characterization, and thermal properties of microencapsulated phase change material for thermal energy storage. Sol Energy Mater Sol Cell. 2009;93:143–7.10.1016/j.solmat.2008.09.009Search in Google Scholar

[125] Sari A, Alkan C, Karaipekli A, Uzun O. Microencapsulated n-octacosane as phase change material for thermal energy storage. Sol Energy. 2009;83:1757–63.10.1016/j.solener.2009.05.008Search in Google Scholar

[126] Alay S, Alkan C, Göde F. Synthesis and characterization of poly(methyl methacrylate)/n-hexadecane microcapsules using different cross-linkers and their application to some fabrics. Thermochim Acta. 2011;518:1–8.10.1016/j.tca.2011.01.014Search in Google Scholar

[127] Alkan C, Sari A, Karaipekli A. Preparation, thermal properties and thermal reliability of microencapsulated n-eicosane as novel phase change material for thermal energy storage. Energy Convers Manag. 2011;52:687–92.10.1016/j.enconman.2010.07.047Search in Google Scholar

[128] Cho W, Kook JW, Lee SM, Koh WG, Kim JH. Modification of heat storage ability and adhesive properties of core/shell structured phase change material nanocapsules. Macromol Res. 2016;24:556–61.10.1007/s13233-016-4075-8Search in Google Scholar

[129] Fortuniak W, Slomkowski S, Chojnowski J, Kurjata J, Tracz A, Mizerska U. Synthesis of a paraffin phase change material microencapsulated in a siloxane polymer. Colloid Polym Sci. 2013;291:725–33.10.1007/s00396-012-2782-zSearch in Google Scholar PubMed PubMed Central

[130] Konuklu Y, Paksoy H. Polystyrene-based caprylic acid microencapsulation for thermal energy storage. Sol Energy Mater Sol Cell. 2017;159:235–42.10.1016/j.solmat.2016.09.016Search in Google Scholar

[131] Sari A, Alkan C, Altintaş A. Preparation, characterization and latent heat thermal energy storage properties of micro-nanoencapsulated fatty acids by polystyrene shell. Appl Therm Eng. 2014;73:1160–8.10.1016/j.applthermaleng.2014.09.005Search in Google Scholar

[132] Sari A, Alkan C, Özcan AN. Synthesis and characterization of micro/nano capsules of PMMA/capric-stearic acid eutectic mixture for low temperature-thermal energy storage in buildings. Energy Build. 2015;90:106–13.10.1016/j.enbuild.2015.01.013Search in Google Scholar

Received: 2019-11-25
Revised: 2019-12-04
Accepted: 2019-12-04
Published Online: 2020-09-21

© 2020 Amende Sivanathan et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 24.4.2024 from https://www.degruyter.com/document/doi/10.1515/ntrev-2020-0067/html
Scroll to top button