Next Article in Journal
Methanol, a Plugin Marine Fuel for Green House Gas Reduction—A Review
Next Article in Special Issue
Effect of Dropping Speed of Reducing Agent on the Preparation of LA/Ag Phase-Change Nanocapsules
Previous Article in Journal
Investigation and Optimization of Integrated Electricity Generation from Wind, Wave, and Solar Energy Sources
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review on Thermal Properties with Influence Factors of Solid–Liquid Organic Phase-Change Micro/Nanocapsules

by
Huanmei Yuan
1,2,3,
Sitong Liu
1,2,
Tonghe Li
1,2,
Liyun Yang
2,
Dehong Li
4,
Hao Bai
1,2,* and
Xiaodong Wang
4
1
State Key Laboratory of Advanced Metallurgy, University of Science and Technology Beijing, 30# Xueyuan Road, Haidian District, Beijing 100083, China
2
School of Metallurgical and Ecological Engineering, University of Science and Technology Beijing, 30# Xueyuan Road, Beijing 100083, China
3
AI Chip Center for Emerging Smart Systems, Building 17W, 17 Science Park West Avenue, Pak Shek Kok, NT, Hong Kong
4
Department of Wood and Forest Sciences, Laval University, Quebec, QC G1V 0A6, Canada
*
Author to whom correspondence should be addressed.
Energies 2024, 17(3), 604; https://doi.org/10.3390/en17030604
Submission received: 4 January 2024 / Revised: 22 January 2024 / Accepted: 25 January 2024 / Published: 26 January 2024
(This article belongs to the Special Issue Modeling Multiphase Flow and Reactive Transport in Porous Media 2024)

Abstract

:
Solid–liquid organic phase-change micro/nanocapsules are potential candidates for energy storage. Recently, significant progress has been made regarding phase-change micro/nanocapsules in terms of their synthesis, properties, and applications. Extensive research has been conducted to enhance their thermal properties, such as thermal storage capacity, thermal conductivity, and thermal reliability. However, factors that influence the thermal properties of micro/nanocapsules have received little attention. This study presents a comprehensive review of phase-change micro/nanocapsules focusing on their thermal properties and their influencing factors. In addition, the thermal properties of the major solid–liquid organic pure phase-change materials are summarized. Furthermore, common micro/nanoencapsulation methods and their influence on the thermal properties were analyzed. Finally, the potential applications of these phase-change micro/nanocapsules were also investigated. This study was devoted to enhancing the thermal properties of micro/nanocapsules, which play a crucial role in their practical applications.

1. Introduction

There is rising interest in developing and implementing practical processes for storing or transporting thermal energy that may contribute to harnessing renewable energy sources such as solar energy, making better use of non-renewable energy sources, and regulating temperature in certain applications [1,2]. This energy can be stored as sensible or latent heat in materials for thermal energy storage (TES) systems [3,4]; in particular, latent heat is more efficient for TES because the material can maintain a constant temperature during energy storage/release [5,6,7].
Phase-change materials (PCMs) are the primary carriers for thermal energy storage in a latent heat TES system because PCMs can store a considerable amount of heat under nearly isothermal conditions. Phase transition, such as solid-to-liquid, enables materials to store thermal energy without significant changes in volume and/or pressure, making solid–liquid PCMs suitable for various applications [3,4,6,8]. Thus, solid–liquid PCMs are the most prevalent thermal energy storage materials. Solid–liquid PCMs typically include organic, inorganic, and eutectic substances [9,10,11,12,13]. Paraffins, fatty acids, and alcohols are among the frequently used solid–liquid organic PCMs. Solid–liquid inorganic PCMs are classified into metals, salts, and hydrates. Eutectics are a combination of two or more chemical compounds, such as capric–palmitic acid. Among these, solid–liquid organic PCMs are the most commonly studied for low-temperature (below 100 °C) applications, as they have the potential for numerous applications [14,15,16,17] such as solar energy to make better use of residual thermal energy or renewable energy sources, food industry and textile production to regulate temperature changes [18,19,20,21,22,23], buildings with refrigeration and heating systems to ensure energy-saving, and thermal management of electronic devices [24,25]. In practical applications, phase-change capsules formed by encapsulation technology outperform those formed directly from PCM because encapsulation can prevent morphological changes during phase transitions, increase the specific surface area with small particles, prevent the loss of liquid PCM, limit the exposure of the core PCM to the environment, and increase thermal conductivity by employing a thermally conductive shell. According to their size, encapsulated PCMs can be classified into macroencapsulated PCMs (>1000 μm), microencapsulated PCMs (1–1000 μm, MEPCMs), and nanoencapsulated PCMs (<1 μm, NEPCMs) [26,27]. There is significant interest in the synthesis of micro/nanocapsules with larger specific surface areas to facilitate heat transfer. These micro/nanocapsules are favored for use in working fluids that can not only store heat but also achieve rapid heat transfer.
Thermal properties, such as thermal storage capability, thermal conductivity, and thermal reliability, are the most important characteristics of phase-change micro/nanocapsules [28]. The thermal storage capability is generally evaluated by the heat stored per unit mass of micro/nanocapsules, also known as latent heat (ΔH). A high ΔH indicates a high thermal storage density, which is desirable for practical applications. Thermal conductivity represents the heat transfer capacity of the micro/nanocapsules and is an important performance indicator for determining whether the micro/nanocapsules can rapidly release or absorb heat. Therefore, thermally conductive micro/nanocapsules are ideal for most applications. The thermal reliability of micro/nanocapsules is a practical performance parameter that represents their service life. Generally, good thermal reliability is expected if there is little loss of ΔH after extensive thermal cycling. Recent reviews have focused on the thermal properties of micro/nanocapsules. For instance, Leong et al. [29] discussed the phase-change temperatures, latent heat, thermal conductivity, and phase-change duration of nanoparticles. Regarding the thermal storage capacity, they indicated the requirement for additional investigation into the primary elements influencing this property. For thermal conductivity, they summarized the factors influencing the thermal conductivity of nano-enhanced PCM. In addition, various researchers [30,31,32,33] have reviewed methods for improving PCM thermal conductivities. Lin et al. [30] compared these methods and concluded that adding thermally conductive fillers was the most effective way to improve the thermal conductivity of PCMs. In the duration section, references were briefly presented to illustrate the changes in the phase-change temperature, thermal conductivity, and thermal storage capacity prior to and following the thermal cycling test. Felipe et al. [34] examined the influence of micro/nanocapsule composition on their thermal properties. They demonstrated that employing appropriate synthetic methods and strategies to control the composition (such as altering the core or shell composition) enabled the preparation of MEPCMs and NEPCMs with better performance.
Notably, based on the aforementioned reviews, these studies rarely involved systematic summaries and analyses of the three key thermal properties of micro/nanocapsules. For example, numerous reviews on thermal conductivity are primarily concerned with a summary of enhancement methods; nonetheless, the factors affecting thermal conductivity must also be analyzed. Various studies have been devoted to developing micro/nanocapsules with good thermal storage capacity; however, the thermal storage capacity is affected by the core latent heat, shell density, shell thickness, particle size, etc. Thus, it is necessary to systematically analyze the factors influencing the thermal storage capacity of micro/nanocapsules. In addition, thermal reliability is an important performance parameter accompanying the preparation of micro/nanocapsules; however, there have been few reviews on the thermal reliability of micro/nanocapsules to date, preventing the evaluation of performance integrity in the future development of capsules. Accordingly, this study aims to gain insight into the thermal properties and factors influencing the utilization of micro/nanocapsules for diverse applications.
The workflow distribution of this study consists of five major sections in addition to the conclusion. In Section 2, the properties of solid–liquid organic PCM and their challenges are introduced; in particular, paraffin and fatty acid-based PCMs are summarized considering their nature and thermal properties. In Section 3, micro/nanoencapsulation of solid–liquid organic PCMs is elaborated with fabrication methods and thermal properties. The size and encapsulation analysis of micro/nanocapsules with different methods, including the preparation trend over the past years, help to understand the development of the ideal thermal performance evaluation with current studies. Section 4 summarizes the research on the thermal properties of solid–liquid organic PCM micro/nanocapsules, including thermal storage capacity, thermal conductivity, and thermal reliability. Section 5 presents the factors that influence these thermal properties. Section 6 introduces the possible applications of solid–liquid organic phase-change micro/nanocapsules.

2. Solid–Liquid Organic PCMs as Core Materials of Micro/Nanocapsules

The selection of the core PCM for micro/nanocapsules is the key to their thermal storage performance. Solid–liquid organic PCMs are suitable for various applications with temperatures below 100 °C [31]. Owing to their excellent properties, such as consistent melting, narrow phase transition temperature range, little or no supercooling, self-nucleation, high thermal energy density, and lack of phase segregation, they are preferred for energy-efficient buildings, solar energy storage, working fluids, thermal management of electronic devices, batteries, etc. [35].

2.1. Paraffin-Based PCMs

As a mixture of n-alkanes with straight chains, paraffin waxes have the general formula CH3–(CH2)n–CH3, which can be classified based on the number of carbon atoms and their physical state. Specifically, n-alkanes with no more than four carbon atoms include pure alkanes in the gas phase, carbon atoms between 17 and 45 are liquid paraffins, and carbon atoms greater than 45 are solid waxes at room temperature.
Paraffin-based PCMs refer to solid paraffin waxes, which are mixtures of saturated hydrocarbons including linear, isomeric, hyperbranched, and naphthenic hydrocarbons [36]. Commercial paraffin waxes contain mixtures of isomers whose melting temperatures vary and increase with the chain length. Paraffin waxes commonly exhibit a high thermal storage capacity, although this value is not proportional to increasing chain length (Table 1). Generally, paraffin waxes with a phase-change temperature within 100 °C are inexpensive, safe, non-irritating, and reliable. Considering the chemical point and economics, most industrial-grade waxes are inactive and stable and can be used in latent heat TES systems for lengthy crystallization-melting cycles owing to their high stability [37]. Table 1 lists certain thermal properties of paraffins as low-to-medium temperature (within 100 °C) thermal storage materials. In addition to its favorable properties, paraffin also exhibits some undesirable properties, such as low thermal conductivity and moderately high flammability [38], which can be partially eliminated using additives or paraffin compounds.

2.2. Fatty Acid

Fatty acids are carboxylic acids composed of a long hydrocarbon chain (possibly saturated or unsaturated) and a terminal carboxyl group. The majority of naturally occurring fatty acids have an even number of carbon atoms and lack branched chains (CH3(CH2)2nCOOH, n (4–28)). Saturated fatty acids lack a C=C double bond [40]. They share the same formula of CH3(CH2)nCOOH, with only a difference in “n”. The most commonly employed saturated fatty acids (SFAs) are lauric acid (LA), palmitic acid (PA), capric acid (CA), stearic acid (SA), and myristic acid (MA). Different fatty acids or fatty acids with other solid–liquid organic PCMs can favorably form eutectic mixtures or synthetic esters, allowing the adjustment of the phase transition temperature.
Fatty acids have gained considerable interest as potential TES candidates among solid–liquid organic PCMs. Around 1989, D. Feldman evaluated the thermal properties of CA, LA, MA, PA, SA, and their binary blends [41]. These fatty acids exhibited a melting temperature range of 30–65 °C and a melting latent heat of 153–182 J/g. Feldman et al. [42] extensively studied the applications and thermal stability of fatty acids. The thermal properties of SA, PA [43], and MA [44] have also been analyzed. Cedeno et al. [45] obtained descriptive data regarding the melting temperature and heat of fusion of PA, SA, oleic acid, and their binary and ternary blends. Gbabode [46] characterized the thermal properties of the crystal components of odd-numbered fatty acids, from tridecanoic acid to tricosanoic acid.
Currently, fatty acids and their derivatives are recognized as attractive PCMs that could serve as energy-retaining compounds in solar energy heat utilization systems and buildings owing to their relatively efficient thermal, physical, and chemical properties and their easy incorporation into composite materials [47]. The properties of fatty acid-based PCMs are summarized in Table 2.

2.3. Other Organic Solid–Liquid PCMs

In addition to paraffins and fatty acids, other types of solid–liquid organic PCMs also exist, including sugar alcohols, esters, and glycols. Sugar alcohols, also called polyalcohols, are regarded as medium temperature PCMs (90–200 °C) and have not received much attention. Hormansdorfer first mentioned that polyalcohols can be used as PCMs [5]. Subsequently, Talja and Roos [50] and Kaizawa et al. [51] studied their phase-change behavior and concluded that certain polyalcohols exhibit a latent heat nearly twice that of other organic PCMs. However, they exhibit a large degree of supercooling, which hinders the TES efficiency. For instance, A. Sari [52,53] indicated that erythritol exhibited a melting/solidifying temperature of 118.4/36.22 °C with a corresponding latent heat of 379.57/255.95 J/g. Evidently, the latent heat value for solidification is approximately 32.6% lower than that of melting, which indicates that only 67.4% of the stored thermal energy could be used. This mismatch leads the TES to deviate from the original path, which is unacceptable for practical applications.
Esters, which are derived from acids by substituting one hydroxyl group (–OH) with an alkyl (–O) group, exhibit a narrow temperature range during the solid–liquid transformation, and eutectics can be formed with minimal or no supercooling. In addition, Nikolic’ et al. [54] used a differential scanning calorimeter (DSC) to examine the solid–liquid phase-change process of five fatty acid esters. The samples included methyl palmitate, methyl stearate, hexadecyl palmitate, hexadecyl stearate, and their eutectic mixtures, and 50 thermal cycles were conducted between temperatures of −10 and 60 °C. The tests were repeated after storing these samples for 18 months, and the results demonstrated no changes in their thermophysical properties.
Polyethylene glycol (PEG, POE, and PEO), which consists of dimethyl ether chains (HO–CH2–(CH2–O–CH2–)n–CH2–OH), has been extensively and experimentally investigated and found to exhibit chemical and thermal stability, nonflammability, non-corrosiveness, nontoxicity, and low price [55,56,57,58]. Notably, its phase-transition temperature can be adjusted by adjusting the molecular weight. Sarier and Onder [59] demonstrated that the melting temperature and latent heat increase with an increase in the molecular weight of PEG. For example, for PEG 400, PEG 2000, and PEG 20000, the melting temperatures were 3.2, 51, and 68.7 °C, respectively, with latent heat values of 91.4, 181.4, and 187.8 J/g, respectively. However, PEG with molecular weights exceeding 10,000 would exhibit decreased latent heat because the PEG crystallinity decreases with long molecular chains. The thermal properties of sugar alcohols, esters, and glycol PCMs are listed in Table 3.

2.4. Challenges of Solid–Liquid Organic PCMs

Over the past years, more and more studies have focused on paraffins, PEG, and fatty acids, owing to their adjustable phase transition temperature and high economic benefits (Figure 1). Among paraffins, n-octadecane, n-eicosane, n-hexaoxane, and n-octadecane are popular medium- and low-temperature PCMs for TES. Meanwhile, SA, PA, and LA are the most commonly studied fatty acid PCMs.
However, the properties of these PCMs may be unsatisfactory. Paraffin has low thermal conductivity and flammability. Its volume fluctuates drastically during the phase change. Fatty acids are unstable at high temperatures and have low thermal conductivities [62]. In addition, the leakage problem of the liquid state restricts the application of these pure PCMs [31].

3. Micro/Nanoencapsulation of Solid–Liquid Organic PCMs

It is widely considered that encapsulating PCMs is an effective means for preventing problems associated with solid–liquid phase transitions. In addition to preventing liquid leakage when combining a PCM with other materials, encapsulation is generally beneficial for protecting the PCM by forming capsules that take the PCM as the core, thereby ensuring that the encapsulated PCM has enhanced thermal reliability, controlled thermal energy release, reduced corrosion, increased heat transfer efficiency, heat transfer area, and protection against degradation during endothermic/exothermic cycling. Therefore, encapsulation is an effective method for promoting the application of solid–liquid PCMs as it employs solid material as a shell that envelops the PCM core so that the phase transition is restricted within the shell, thereby maintaining the solid state of the particles [63].

3.1. Encapsulation Methods of Solid–Liquid Organic PCMs

Microencapsulation produces capsules with a size below 1000 μm, which are called microcapsules, and capsules with a size below 1000 nm are regarded as nanocapsules. To date, the most commonly used microencapsulation methods include in situ polymerization, interfacial polymerization, suspension polymerization, emulsion polymerization, miniemulsion polymerization, and sol–gel [64]. Based on these methods, the capsule sizes have evolved from microcapsules to nanocapsules, and the organic shell materials have been replaced by inorganic ones.

3.1.1. In Situ Polymerization

In situ polymerization is a PCM encapsulation method. This process begins by mixing two immiscible liquids, such as water and an organic solvent that contains complementary and direct-acting organic intermediates, which then mutually react to form solid pre-condensates [65]. Melamine–formaldehyde (MF) and urea–formaldehyde (UF) are two widely used shell materials because of their chemical stability and high mechanical strength. Figure 2 depicts a schematic of the in-situ polymerization process [66]. Using the situ polymerization process, Rao et al. [67] prepared an n-docosane/melamine resin MEPCM with particle sizes of 5–50 μm. At a stirring rate of 6000 rpm, the prepared MEPCM with a particle size of 10 μm exhibited an encapsulation ratio of 60% and a melting latent heat of 150 J/g.
Nanocapsules can also be synthesized using an in situ polymerization method. Du et al. [68] fabricated NEPCMs with n-octadecane as the core and phosphorus–nitrogen diamine (PNDA)-modified MF as the shell via in situ polymerization. During this process, the PNDA-modified MF prepolymer was carefully dropped into the previously prepared n-octadecane emulsion with continuous stirring. After filtering, the prepared sample was washed using a 50 wt.% aqueous ethanol solution to remove the impurities, and unencapsulated n-octadecane flame-retardant NEPCMs were obtained after drying in a vacuum environment for 24 h. The diameters of the prepared NEPCMs ranged from 80 to 140 nm, and their latent heats of fusion ranged from 110.8 to 141.3 J/g.
In situ polymerization is widely used to prepare micro-/nanoencapsulated PCMs. As reported [69], the in situ method can produce the best quality microcapsules in terms of the diffusion tightness of their shells and a size range of 5–100 μm. This method is beneficial owing to its cost-effectiveness, ease of automation, and compatibility with the environment. However, residual formaldehyde may be unavoidable after shell formation, leading to environmental and health problems during polymerization [70].

3.1.2. Interfacial Polymerization

The interfacial polymerization method involves dispersing the organic phase (including multifunctional monomers and/or oligomers) together with the core material, which is encapsulated in an aqueous phase [71]. The interfacial polymerization method can also be employed to prepare microcapsules and nanocapsules. Figure 3 illustrates a schematic of micro/nanocapsules formed by interfacial polymerization [72]. Zou et al. [73] fabricated hexadecane/polyurea microcapsules via the interfacial polymerization method; the microcapsules with a size of 2.5 μm possessed a latent heat of 66 J/g at a melting point of 18 °C. Shi et al. [72] devised a one-step interfacial polymerization method to prepare paraffin/polymethyl methacrylate (PMMA) nanocapsules. The results indicated that the nanocapsules with particle sizes ranging from 200 to 400 nm had a spherical shape with a homogeneous, smooth, and compact surface.
Interfacial polymerization is commonly employed to synthesize microcapsules. Commercially produced microcapsules exhibit sizes of 20–30 μm and 3–6 μm, where the former are employed in pesticidal and herbicidal applications and the latter for carbonless paper ink [69]. This method is advantageous as it delivers a high encapsulation ratio and can protect active substances. However, the encapsulated material must be acid- and alkali-resistant to react with the monomer, and it also exhibits the problem of dealing with excess monomers [70].

3.1.3. Suspension Polymerization

The suspension polymerization method involves suspending the water-immiscible reaction mixture as droplets in an aqueous continuous phase [74]. These droplets were formed on vigorously stirring the mixture, and the stabilizer was dissolved in the aqueous phase. Subsequently, polymerization was initiated at the desired temperature until completion. Figure 4 depicts a schematic of the formation process of the suspension polymerization method. The suspension polymerization method can be employed for both microcapsules and nanocapsules. You et al. [75] fabricated n-octadecane/styrene (St)–divinylbenzene (DVB) copolymer microcapsules via suspension polymerization. The microcapsules exhibited an average diameter of 80 μm and a melting fusion of 126 J/g. Zhang et al. [76] used the suspension polymerization method to fabricate n-octadecane/poly(2,3,4,5,6-pentafluorostyrene) nanocapsules, and results demonstrated that nanocapsules with an average size of 490 ± 16 nm exhibited good morphology, high latent enthalpy (171.8 ± 2.2 J/g), and excellent thermal stability.
This method is advantageous because it has a high encapsulation efficiency, the prepared capsules exhibit a smaller particle size, and the process is smooth. However, auto-acceleration is unavoidable in this method, and dispersants are required during polymerization [70].

3.1.4. Emulsion Polymerization

In emulsion polymerization, the initiator should first be dissolved in the aqueous phase; monomers should be emulsified in the polymerization medium and distributed among the droplets, surfactant micelles, and aqueous phase by employing surfactants, as depicted in Figure 5. The commencement of the polymerization reaction lies in the aqueous phase because the initiator can only exist in this phase, which later extends to the micelles. The proportion of monomer molecules dissolved in the aqueous phase is crucial to the size of the obtained particles. In addition, the concentration of the emulsifier, initiator, and polymerization temperature also influence the size [78]. Emulsion polymerization is viable for both microcapsules and nanocapsules. Sari et al. [79] adapted an emulsion polymerization method to fabricate micro/nanocapsules with an MA–PA eutectic mixture as the hybrid core and PMMA as the shell. Microcapsules with particle sizes of 0.1–70 μm possess a latent heat of 68–100 J/g at the phase-change temperature. Using the same method, Baek et al. [80] prepared n-hexadecane nanocapsules coated with the poly(glycidyl methacrylate) macromolecule poly(methylmethacrylate-co-glycidyl methacrylate). The size of the prepared nanocapsules was 260 nm, with the melting/solidifying temperatures and latent heats being 17.23/14.85 °C and 148.05/147.63 J/g, respectively; thus, the encapsulation ratio was 62.46%.
This method proves beneficial as it possesses a fast polymerization rate, a high molecular weight, and allows direct use of an emulsion. However, the polymer separation process is complex, and the product involves large quantities of residual impurities [70].

3.1.5. Miniemulsion Polymerization

In this technique, polymerization occurs within the tiny droplets, which are dispersed while maintaining their size at the nanoscale under the action of large shear forces, usually involving emulsifiers, monomers, water, and initiators [81]. A schematic of the mini-emulsion polymerization process is shown in Figure 6 [77]. Miniemulsion polymerization is a common technique employed to produce nanocapsules. Rezvanpour et al. [82] used the mini-emulsion polymerization process to prepare nanocapsules where n-eicosane served as the PCM core and PMMA served as the shell. The mean diameter of the n-eicosane/PMMA nanocapsules prepared was 135 nm, and under a weight ratio of 50/50, the melting/solidifying temperatures were 34.66/32.92 °C, and the corresponding latent heats were 124.7/119.13 J/g, respectively.
Presently, the common shell materials for nanocapsules prepared using this method are styrene (St)-methylmethacrylate copolymer, polyurea, polystyrene (PS), PMMA, and styrene-butyl acrylate (SBA) [81]. This method is advantageous because the reaction temperature is stable and does not involve volatile solvents. However, it requires a high-energy mixing process [70,83].

3.1.6. Sol-Gel

The sol-gel method [70] is a wet chemical process for producing inorganic oxide materials. Highly chemically active precursors are used to obtain a stable and transparent solution via hydrolysis and condensation, and a shell can be obtained following the gel process. A schematic illustration of the formation process via the sol-gel method is depicted in Figure 7 [84]. The sol-gel method can be employed to prepare both microcapsules and nanocapsules. Wang et al. [85] studied the sol-gel method to produce a silica shell by combining the oil/water (O/W) emulsion technique for the first time, and the resulting spherical microcapsules exhibited a size range of 4–8 μm. Fang et al. [86] and Zhang et al. [87] also prepared the microcapsules using paraffin as the core and SiO2 as the shell by employing the sol-gel method and obtained an encapsulation ratio exceeding 85%. Later, Latibari et al. [88] fabricated nanocapsules using the sol-gel method with PA as the core and SiO2 as the shell. By adjusting the solvent pH value from 11 to 12, the size was enhanced from 183.7 to 722.5 nm, and the encapsulation ratio was enhanced from 83.25 to 89.55%. Subsequently, they synthesized SA/TiO2 (titania) nanocapsules using the same method [89], and the prepared nanocapsules exhibited sizes of 583.4–946.4 nm and presented an encapsulation ratio of 30.36–64.76%.
The sol-gel process is a method for preparing micro/nanocapsules with inorganic shells such as SiO2, TiO2, and ZrO2 [90]. This method is advantageous because it possesses a high reaction efficiency and the mixing process is uniform. However, the reaction is time-consuming, and the raw material is relatively expensive [70].

3.2. Analysis of the Encapsulation Methods of Solid–Liquid Organic PCMs

The synthesis method is pivotal for obtaining micro/nanocapsules with good thermal properties because it determines the shell material, size, and properties of the capsules. The methods of interfacial, in situ, emulsion, suspension, and miniemulsion polymerizations are employed to prepare micro/nanocapsules with organic shells (such as PMMA), whereas the sol–gel method is employed to prepare micro/nanocapsules with inorganic shells (like SiO2, TiO2, etc.).
The major disadvantage of organic shells is their extremely low thermal conductivity, which retards heat exchange with the environment during heat absorption/release cycles, supercooling, and overheating [6]. Compared to organic shells, inorganic shell materials exhibit a relatively high thermal conductivity, which renders them favorable for micro/nanocapsule applications requiring fast heat transfer. Although the inorganic shells could improve the thermal conductivity, their enhancing effect was insignificant owing to their moderate thermal conductivity. Laghari et al. [91] used TiO2 as the shell material to obtain a paraffin-based nano-enhanced PCM with improved thermal conductivity, and the results indicated that using TiO2 (4.8–11.8 W/(m·K)) improved the thermal conductivity of the paraffin-based nano-enhanced PCM from 0.2 W/(m·K) to 0.3690 W/(m·K). However, a minor improvement in thermal conductivity does not significantly enhance their application utility. Therefore, further research is necessary to enhance the thermal conductivity.
In addition to the shell material, the size and encapsulation of the micro/nanocapsules produced using different methods over the previous decades were analyzed (as shown in Figure 8a). The methods of interfacial polymerization, in situ polymerization, emulsion polymerization, suspension polymerization, and sol-gel involve phase-change capsules from nano to micro size, whereas the miniemulsion polymerization method is the only method for preparing nanocapsules. In situ polymerization is predominantly used to prepare microcapsules, and the sol-gel method focuses on preparing nanocapsules between 100 and 1000 nm. The latent heat and encapsulation ratio (Figure 8b and Figure 8c, respectively) of micro/nanocapsules prepared using different methods can all achieve high values because the performance can be adjusted by varying the preparation conditions. Recently, other synthesis methods have gradually occupied a larger proportion of micro-nanocapsules, and related research has dropped drastically in the previous year (Figure 8d and Figure 8e, respectively), indicating that researchers are working on expanding the research of shell materials to further strengthen the performance of micro/nanocapsules, such as using highly thermally conductive metal shell materials to enhance the thermal conductivity of micro/nanocapsules [92].

4. Thermal Properties of Solid–Liquid Organic Phase-Change Micro/Nanocapsules

In applications involving encapsulation, the important aspects of the performance of encapsulated phase-change materials (EPCMs) include thermal storage properties, thermal conductivity (λ), and thermal reliability. Generally, thermal storage properties are evaluated using the ΔH. The following equations are commonly used to analyze the thermal storage properties:
R = H m , E P C M s H m , P C M s × 100 %    
E = H m ,   E P C M s + H c , E P C M s H m , P C M s + H c , P C M s × 100 %
ΔHm,PCMs, ΔHc,PCMs, ΔHm, EPCMs, and ΔHc, EPCMs indicate the melting latent heats of pure PCMs, solidifying latent heats of pure PCMs, melting latent heats of EPCMs, and solidifying latent heats of EPCMs, respectively. R represents the encapsulation ratio of the core materials for the EPCMs, and E reflects the encapsulation efficiency of the PCM. R and E were both positively associated with the core/shell mass ratio. R is typically employed for evaluating EPCM encapsulation owing to the fact that the melting and solidifying latent heats often exhibit minor differences for the solid–liquid EPCMs. The thermal storage properties of the EPCMs were analyzed using a DSC device. Thermal conductivity is an important thermal performance parameter for EPCM applications, particularly for applications in a working fluid, as it affects the heat transfer of the TES system. Typically, EPCM thermal conductivity can be tested using a thermal conductivity analyzer with a laser flash or steady-state plate method. Thermal reliability is also crucial for practical applications, as it could evaluate the EPCM performance after long-term service. Typically, thermal reliability is tested via thermal cycling by heating/cooling the as-prepared samples for a large number of cycles.

4.1. Thermal Storage Properties Overview of Various Micro/Nanocapsules

Thermal storage properties are key performance metrics for micro/nanocapsules applications. A large thermal storage capacity and encapsulation ratio are the expected objectives of EPCM synthesis, because EPCMs with a high thermal storage capacity can absorb and release a large amount of heat in a smaller mass, which is highly efficient for applications. Early research was primarily devoted to the preparation of phase-change microcapsules and their thermal storage properties. Sari et al. [93] prepared the microencapsulated CA, LA, and MA with polystyrene (PS) by employing the emulsion polymerization method, and they obtained microcapsules with mean diameters of 7.7–42.0 μm and latent heats of 87–98 J/g for melting and 84–96 J/g for solidifying, corresponding to the R range of 43.56–48.7%. Moreover, the melting and solidifying phase-change temperature ranges are 22–48 °C and 19–49 °C, respectively. Later, they [94] fabricated n-nonadecane/PMMA microcapsules using the same method, and microcapsules were produced with a mean diameter of 8.18 μm and melted at 31.23 °C, with the latent heat increasing to 139.20 J/g and the R increasing to 60.3%. Further, they [95] synthesized a series of micro/nanocapsules (0.01–115 μm) with a eutectic mixture of tetracosane (C24)-octadecane (C18) as the core and PS as the shell using an emulsion polymerization process. With a shell/core ratio of 2:1, 1:1, and 1:2, the melting latent heats of the PS/(C24-C18) micro/nanocapsules were observed as 71.73, 116.32, and 156.39 J/g, respectively, which correspond to the encapsulation ratio of 29.6, 48.0, and 64.4%, respectively. Figure 9 depicts the morphology and DSC results for the fatty acid/PS, n-nonadecane/PMMA, and C24-C18 eutectic mixture/PS microcapsules.
Zhang et al. [96] employed a suspension-like polymerization method to obtain paraffin wax/PMMA microcapsules and micro/nanocapsules, both of which have a high core content. By increasing the core/shell ratio, the paraffin contents in microcapsules (94 μm) could approach 89.5% with a corresponding latent heat of 137.2 J/g, and for the prepared micro/nanocapsules (0.1–19 μm), the paraffin contents approached 80.2% with a latent heat of 123.0 J/g.
Using an in-situ polymerization method, Yu et al. [97] employed an M/F resin shell to synthesize n-dodecanol microcapsules. The obtained microcapsules (30.6 μm in diameter) melt at 21.5 °C, and the latent heat and R of the microcapsules were as high as 187.5 J/g and 93.1%, respectively, when the emulsifier/n-dodecanol ratio was 4.8%.
Rao et al. [67] prepared the n-docosane (C22H46)/melamine resin microcapsules via the in-situ polymerization method. The prepared microcapsules (approximately 10 μm) exhibited an R of 60% and demonstrated a high melting latent heat of 150 J/g.
Fang et al. [98] fabricated n-tetradecane/UF resin nanocapsules (average particle size of 100 nm) via in situ polymerization. By increasing the C6H6O2 (resorcin, modifier) concentration from 0.25 to 5%, the R increased from 30.4 to 61.8%, and the melting and crystallization latent heat increased from 66.01 and 66.39 J/g to 134.16 and 134.5 J/g, respectively.
Existing studies show that encapsulation ratio and phase-change enthalpy are commonly used parameters to characterize thermal energy storage performance. Besides, the volumetric energy storage density, ΔHv, may also be used to characterize thermal storage capacities as it also considers the density of phase-change micro/nanocapsules, which is an important property for the micro/nanocapsules.
Therefore, we derived the ΔHv based on the enthalpy ΔHm, encapsulation ratio R, and density ρ of EPCMs.
Δ H V = ρ H m , P C M s    
ρ = ρ s h e l l 1 R V + ρ c o r e R V
R V = 1 1 + ρ c o r e ρ s h e l l ( 1 R R )    
where ΔHv indicates the volumetric energy storage density, (J/g), ρ indicates the density of phase-change micro/nanocapsules, Rv indicates the volumetric encapsulation ratio of the micro/nanocapsules, and ρshell and ρcore indicate the density of the shell and core of the micro/nanocapsules, respectively.
With the research mentioned above, we calculated the ΔHv of the corresponding micro/nanocapsules. The results are shown in Table 4. Due to the density, the calculated ΔHv is relatively smaller than ΔHm, PCMs. This is because of the small density of the core material and shell material. If the micro/nanocapsules show a large density, the value of ΔHv should be increased accordingly, which means that the micro/nanocapsules present a good energy storage density per unit volume. Details are discussed in Section 5.1.4. Since most of the micro/nanocapsules studied show a relatively small density (less than 1 g/cm3), we still use phase-change enthalpies and R to characterize the thermal storage capacity.
To date, an increasing number of micro/nanocapsules exhibiting good thermal storage capacities have been developed. Table 5 presents detailed information on the micro/nanocapsules.

4.2. Thermal Conductivity Overview of Solid–Liquid Micro/Nanocapsules

Thermal conductivity (λ) is an important property for phase-change micro/nanocapsules in various applications because λ affects the heat transfer during the heat absorption and release processes [139]. Generally, a larger λ improves the charge/discharge rate during energy storage/release. For heat dissipation applications in electronic devices, a higher λ is necessary to eliminate excess heat as rapidly as possible.

4.2.1. Thermal Conductivity Enhancement by Additives

Most solid–liquid organic PCMs possess extremely low λ, and the shell materials employed to encapsulate the PCM core predominantly include organic materials, such as PMF resin [140], PUF resin [141], polyurethane (PU) [142], PS [143], and PMMA [144], which also suffer from low λ. Both of these factors limit the practical applications of micro/nanocapsules in heat dissipation systems. Therefore, the λ of the micro/nanocapsules must be improved to achieve a fast heat transfer rate. Several studies have been conducted to overcome the issue of low λ of micro/nanocapsules, and the major techniques adopted to enhance the λ of micro/nanocapsules have been embedding/grafting thermally conductive inorganic materials into the shell. Thus far, numerous studies on improving the λ of micro/nanocapsules have been reported [30]. Jiang et al. [145] synthesized microcapsules with a PMMA shell. After the shell was inlayed by nano-Al2O3, the modified microcapsules (16 wt.% nano-Al2O3) exhibited a higher λ of 0.314 W/(m·K) than the unmodified ones (approximately 0.243 W/(m·K)). Yang et al. [144] prepared n-octadecane microcapsules with an PMMA shell, and silicon nitride (Si3N4) was employed to modify the shell, as depicted in Figure 10. The results indicated that λ was significantly enhanced by adding the highly thermally conductive Si3N4. Wang et al. [146] prepared n-octadecane microcapsules containing PMF and shells. The addition of 7% nano-SiC could increase the thermal conductivity of the PMF–SiC microcapsules by 60.34% compared to that of PMF microcapsules. Further detailed thermal conductivity data with related references are presented in Table 6.

4.2.2. Thermal Conductivity Enhancement by Inorganic Shells

To improve the λ of micro/nanocapsules, various researchers have focused on encapsulation using inorganic shells. Zhang et al. [87] fabricated n-octadecane microcapsules using an inorganic SiO2 shell to improve λ via a sol-gel process. Tetraethyl orthosilicate (TEOS) was used as the SiO2 precursor. When the n-octadecane/TEOS mass ratio was set to 50/50, the λ of the microcapsules was enhanced to 0.6213 W/(m·K). Eventually, the λ of the microcapsules improved significantly compared with that of the pure PCM, owing to the thermally conductive SiO2 shell. Yu et al. [150] synthesized n-octadecane/calcium carbonate (CaCO3) microcapsules using a self-assembly method. The n-octadecane/CaCO3 with a mass ratio of 30/70 exhibited a remarkable improvement in λ, which equaled 1.674 W/(m·K), and correspondingly, the enhancement ratio approached 726%. Chai et al. [151] fabricated n-eicosane/TiO2 microcapsules using a sol-gel process. During the process, tetra butyl titanite (TBT) was employed as the precursor of TiO2. With an n-eicosane/TBT weight ratio of 40/60, the microcapsules exhibited a λ of 0.865 W/(m·K) corresponding to an enhancement of 18.6% compared to n-eicosane. To further improve the λ of micro/nanocapsules, some researchers have focused on integrating thermally conductive fillers into inorganic shells. Zhu et al. [134] prepared n-octadecane/silica nanocapsules, and by grafting Ag nanoparticles into the silica shell (as depicted in Figure 11), the apparent λ of the nanocapsules increased significantly from 0.246 to 1.346 W/(m·K) on changing the AgNO3 concentration from 8 to 48 g/L. Yuan et al. [152] employed graphene oxide (GO) to modify paraffin/SiO2 microcapsules, and they discovered that the λ increased from 1.03 to 1.162 W/(m·K). They also dispersed the two types of microcapsules in a water fluid and determined that the λ of the slurry increased from 0.703 W/(m·K) containing 10 wt.% paraffin/SiO2 microcapsules to 0.713 W/(m·K) containing 10 wt.% paraffin/SiO2–GO microcapsules.

4.2.3. Thermal Conductivity Enhancement by Metal Shell

To further improve the thermal conductivity of micro/nanocapsules, a method was developed in a recent study to fabricate phase-change micro/nanocapsules employing a continuous metal (Ag) shell coating [92] (Figure 12). In this process, AgBr ultrafine solid particles were introduced as emulsifiers to improve the rigidity of the emulsion droplet during the solid-liquid transition, and hydroquinone was subsequently used as a reducing agent to adhere to the emulsion after repeated reduction. The AgBr on the droplet surface eventually formed an Ag shell. Palmitic acid (PA) was employed as the core material. The PA/Ag nanocapsules (100–600 nm in diameter) exhibited excellent thermal conductivity up to 4.072 W/(m·K), indicating that the Ag shell significantly improved the thermal conductivity of the nanocapsules.

4.3. Thermal Reliability Overview of Solid–Liquid Micro/Nanocapsules

In practical applications, PCMs or micro/nanoencapsulated PCMs must possess the ability to store and release heat for numerous cycles without significant changes in the latent heat and phase-change temperatures. Therefore, thermal reliability is crucial for a wide range of applications, and particular studies have also considered this aspect. Abhat [155] considered PCMs within the temperature range of 0–120 °C and indicated that studies of thermal reliability are necessary to evaluate their long-term and short-term usage. Salunkhe and Krishna [156] indicated that PCMs should satisfy both primary and secondary criteria before they are considered for applications and that thermal reliability is crucial for efficient latent heat storage usage.
Thermal reliability is often tested using thermal cycling tests, where the core PCM is heated to a liquid state at a temperature above the melting point and then cooled down to a solid state at a temperature below the freezing point, which represents one cycle for the test. Following a large number of cycles, the samples were tested using DSC to compare the changes in the phase transition temperature and latent heat. Typically, if a micro/nanocapsule can withstand numerous thermal cycles (N) with only a minor change in the heat storage performance, particularly the latent heat, after thermal cycling, this indicates that the micro/nanocapsule possesses good thermal reliability. Several review articles have focused on the thermal reliability of PCMs. Sharma and Sagara [157] presented a list of 250 PCMs considering their melting temperature and latent heat, and they discussed long-term stability studies related to the thermal cycling of these particular PCMs. Rathod and Banerjee [4] presented a detailed review of the thermal reliability of latent-heat-storage materials. They discovered that paraffins were stable after several heating/cooling cycles. They also recommended investigating fatty acids in relation to thermal cycling. Ferrer et al. [158] reported that common standards for performing thermal reliability tests are unavailable.
In addition, various studies on organic solid–liquid phase-change micro/nanocapsules also conducted thermal reliability tests. Thermally enhanced n-octadecan/SiO2–BN nanocapsules were prepared by Lan et al. [159] using a miniemulsion polymerization method. These nanocapsules underwent 500 heating/cooling cycles in the thermal reliability assessment experiment. By analyzing the latent heat and phase-transition temperature changes of the nanocapsules under different cycles, the results indicate that the phase-change temperature fluctuation was within 1.5 °C, implying a minor change in the phase-change temperature following the thermal cycling test. In addition, the latent heat of n-octadecane/SiO2-BN nanocapsules exhibited a variation below 4% after 500 thermal cycles, which nearly remained constant. Zhang et al. [76] produced n-octadecane/poly(2,3,4,5,6-pentafluorostyrene) phase-change nanocapsules with a diameter of 490 ± 16 nm using the suspension polymerization method, and little change in latent heat was observed after 200 thermal cycles for the nanocapsules. Nikpourian et al. [160] prepared paraffin wax/PU nanocapsules and evaluated their thermal reliability via thermal cycling tests, as shown in Figure 13. The results indicated no obvious leakage outside the nanocapsules following 50 and 100 thermal cycles. Detailed thermal reliability data with related references are presented in Table 7.
From the aforementioned studies, although substantial research involved evaluating the thermal reliability using thermal cycling tests, the number of thermal cycling tests in any particular research does not indicate the maximum number of cycles that the prepared micro/nanocapsules can undergo. In addition, the results clearly indicate the absence of a general standard that dictates the number of cycles the micro/nanocapsules should undergo. However, as reported on a daily basis [168], the latent heat TES will experience heat absorption at least once during the daytime and heat release once during the night. Consequently, for each day, there would be a thermal cycle, and within one year, 365 thermal cycles would be performed by the latent heat TES system. Therefore, 1200, 3650, and 10,000 thermal cycles correspond to operational performances equivalent to approximately 4, 10, and 30 years, respectively. This condition can be simulated using an accelerated thermal cycling test in the laboratory, which can be performed by repeating melting/solidification cycles for the PCM under controlled conditions, which would reproduce the long-term performance of the TES system within a relatively short time.

5. Influence Factors on the Thermal Properties of Micro/Nanocapsules

5.1. Influence Factors on the Thermal Storage Properties

For micro/nanocapsules, their morphological, physico-chemical, and thermal properties rely on various synthetic parameters, such as pH, mixing rate, content and type of core PCM, emulsifier, core/shell ratio, reaction time, and temperature [6]. It is difficult to predict the most appropriate preparation recipe for a particular system, and several experiments are required to produce the micro/nanocapsules with highly favorable properties. Among the various synthetic conditions, a few major factors influence the thermal storage properties of micro/nanocapsules, which are summarized in this study.

5.1.1. Mass Ratio of Core/Shell

The mass ratio of the core/shell materials added during the preparation process is a key factor that affects the thermal storage properties. Generally, under suitable preparation conditions, if the additional core material is used, high thermal storage properties and a large encapsulation ratio of micro/nanocapsules can be achieved. Recently, Hu et al. [169] prepared methyl laurate/PU microcapsules using an interfacial polymerization method. They studied the preparation conditions of the microcapsules and discovered that by adjusting the mass ratio of core/shell from 1:1 to 4:1, the melting enthalpies, freezing enthalpies, and encapsulation ratio of microcapsules increased from 79.03 J/g, 70.63 J/g, and 40.71% to 155.6 J/g, 136.4 J/g, and 85.28% (as shown in Figure 14), respectively. Latibari et al. [89] fabricated SA/TiO2 nanocapsules via a sol-gel process. During the process, titanium tetraisopropoxide (TTIP) was used as the precursor of TiO2, and with different SA/TTIP ratios of 50/50, 60/40, 70/30, and 80/20, the obtained nanocapsules exhibited an increased melting latent heat of 58.12, 87.95, 99.02, and 123.96 J/g, respectively, and an encapsulation ratio of 30.36, 45.94, 51.73, and 64.76%, respectively (Figure 15).

5.1.2. pH

In reality, synthesis conditions with a large core/shell mass ratio do not always produce micro/nanocapsules with high latent heat and encapsulation ratios. Liu et al. [170] fabricated n-eicosane/PF microcapsules using an in situ polymerization method and studied the effect of the core/shell mass ratio on thermal storage (Figure 16). With the addition of n-eicosane of 1, 2, 3, 4, and 5 g, the latent heat of the prepared microcapsules changed to 94.86, 99.50, 173.83, 143.94, and 155.36 J/g, respectively, which indicated that the latent heats of microcapsules did not increase with the core/shell mass ratio, and other factors also affected the thermal storage of the microcapsules. Their study remarked that during the fabrication process, phenol and formaldehyde monomers could be polycondensed directly into a solid layer in the polycondensation process instead of forming a shell on the surface of n-eicosane emulsion droplets, resulting in a reduction in the actual amount of n-eicosane within the microcapsules. Therefore, shell-forming conditions are also important for thermal storage capacities. In our previous study [138], the effect of alkaline pH on LA/SiO2 nanocapsule preparation and thermal storage capacities was explored. We had fixed the core/shell ratio at a certain value and adjusted the pH of the shell precursor solution in the sol-gel process to study the formation of nanocapsules and thermal storage properties. Eight samples were prepared and investigated. The results demonstrated that the reasonable pH range for nanocapsule formation lies between 9.9 and 10.9, within which the nanocapsules obtained presented an encapsulation ratio as high as 83.0%, and correspondingly, the latent heats approached 160.0 J/g, implying a large thermal storage capacity. When the pH was lower than 9.9, the nanocapsules exhibited low latent heat (less than 100 J/g), whereas when the pH was higher than 10.9, nanocapsules with core-shell structures could not be produced. Therefore, controlling shell formation is also crucial for the heat storage capacity of nanocapsules, which is predominantly due to its influence on the shell thickness or particle distribution of the capsules.

5.1.3. Emulsion Droplet Size

Furthermore, the formation condition of the core is also highly important for the thermal storage capacities. Theoretically, the micro/nanocapsule size relies on the core PCM emulsion droplets, which form a solid-liquid organic PCM that can be dispersed in the solution and subsequently coated with thin shell materials. Large emulsion droplets tend to produce large micro/nanocapsules with high latent heat because large cores may occupy high mass ratios in the final micro/nanocapsules. However, the latent heat may also be affected by other factors during the preparation process. In one of our studies [137], we studied the effect of particle size on thermal storage capacity. We prepared LA/SiO2 nanocapsules under different conditions; these nanocapsules exhibited varying latent heats with their corresponding mean diameters (Figure 17). On analyzing the relation of the latent heat with particle size, it was observed that an optimum particle size of 340 nm in diameter corresponds to the highest latent heat. For particle sizes below 340 nm, the latent heat increased with the increase in particle size. In contrast, for particle sizes exceeding 340 nm, the latent heat exhibited a decreasing trend with increasing particle size for the nanocapsules. This is primarily related to the relationship between the cumulative percentage of droplets and the generation rate of silica clusters. When the cumulative percentage of the droplets increased faster than the generation of the corresponding clusters, the shell tended to be thinner and the core tended to be larger, which led to an increased thermal storage capacity with particle size and encapsulation ratio (zone I in Figure 18a). When the clusters are generated faster than the increase rate of the core particle size, the shell thickening results in a decreased thermal storage capacity with particle size. When the cumulative percentage of the droplet size is gradually balanced by that of the cluster generation rate, the relative ratio of the core to shell attains a maximum value, and thus the optimum particle size, corresponding to the highest latent heat, is obtained.

5.1.4. Summary of the Influence Factors on Thermal Storage Capacity

From the aforementioned analysis, it is observed that regardless of the factors affecting the latent heat of micro/nanocapsules, they all contribute to their final core/shell mass ratio. Generally, micro/nanocapsules with larger cores and thinner shells exhibit good thermal storage capacity. This is an inference based on the premise that the shell material density is relatively small (such as SiO2, approximately 2.2 g/cm3). If the shell density significantly exceeds that of the core (such as Ag shell, 10.5 g/cm3), the measured latent heat of the nanocapsules will still be small, primarily owing to the larger mass fraction of the shell material occupied within the micro/nanocapsules. Theoretically, the latent heat is obtained according to the mass fraction of the capsule occupied by the core with the identical amount of core PCM; the latent heat calculated using the shell of a small density is relatively higher than that of a large density, which creates the illusion that the micro/nanocapsules with heavy shells show poor thermal storage capacity (as shown in Figure 19a), whereas the energy stored in the core is the same. Thus, the evaluation by latent heat is not appropriate, and a new evaluation parameter should be derived. RV, defined as the volume ratio of the core occupying the whole micro/nanocapsule, was proposed in our previous study [92] to evaluate the encapsulation efficiency of the micro/nanocapsules, considering R and the difference in density (k = ρ c o r e ρ s h e l l ) between the core and shell materials. Figure 19b depicts a comparison of R and Rv for different k levels. From the structural comparison, we observe that different k values lead to the visual miscalculation of R, whereas RV can eliminate this miscalculation to present the real core-shell structure state. Nevertheless, micro/nanocapsules having a ‘light’ shell are the first choice for the application.

5.2. Influence Factors on Thermal Conductivity

In the present study, the methods to enhance the thermal conductivity of micro/nanocapsules mainly included embedding/grafting thermally conductive particles into the shell and encapsulating PCMs with high thermal conductivity shells. Both methods could effectively improve the λ of the phase-change micro/nanocapsules, whereas some influencing factors are important in the enhancement.
For thermally conductive particles and enhancement of the micro/nanocapsules, sufficient addition of nanoparticles is required to develop a network that is conducive to thermal conduction [171], which is supported by Zhang et al. [172] in their experimental study on the thermal properties of short carbon fiber/erythritol PCMs. Based on this, the λ of the phase-change micro/nanocapsules relies on the particle concentration, particle dispersion, and interface interaction between the particle and shell.

5.2.1. Thermal Conductive Nanoparticle Content

Various studies have reported that the λ of enhanced micro/nanocapsules increases with increasing nanoparticle content [134,144,145,146,147,148,149]. Li et al. [149] added up to 4 wt.% of carbon nanotubes to the urea formaldehyde resin (UFR) shell of paraffin/UFR microcapsules, and an increase of λ from 0.095 W/(m·K) (1 wt.% GO) to 0.146 W/(m·K) (4 wt.% GO) was observed. Park et al. [148] studied the fabrication of paraffin/polyurea-(nano Fe3O4) nanocapsules. They confirmed that the λ of nanocapsules with 6.6 w.t% nano Fe3O4 increased by 155% (0.342 W/(m·K)). Chen et al. [147] studied the λ of GO-enhanced microcapsules by increasing their GO content. When the GO content was below 1 wt.%, an increase in λ was obtained from 0.1157 to 0.1738 W/(m·K). When the GO content was 1 wt.%, the λ of the GO-enhanced microcapsules increased by 50.21%. When the GO content was 4 wt.%, the λ of the GO-enhanced microcapsules was enhanced by 66.29%. For this enhancement, the oxygen-containing functional groups on the GO surface enabled GO to disperse well in the MF prepolymers, and the GO sheet network had a positive effect on the λ of the GO-enhanced microcapsules. Hence, a minor amount of GO can significantly improve the λ of the GO-enhanced microcapsules. However, the networks became saturated as the GO content increased to 4 wt.%; thus, there was an insignificant increase at the end.

5.2.2. The Distribution and Interface Interaction of Particles in the Shell

The λ of micro/nanocapsules with high thermal conductivity nanoparticles added to the shell cannot always exhibit larger values than those with low thermal conductivity nanoparticles [29]. The dispersion state of the highly conductive nanoparticles (fillers) in the shell and the interfacial interaction between the nanoparticles (fillers) and shell majorly influence the λ of micro/nanocapsules. The former is related to the thermal resistance of the interface, and the latter involves the formation of a thermal conduction path.
Liu et al. [173] prepared dodecanol/MF resin microcapsules and modified the shell with a GO–CNT hybrid filler to improve λ. Herein, by adding GO and CNTs, the λ of the microcapsules could be enhanced, and the GO–CNT hybrid filler was preferable to that of individual GO or CNT. Specifically, the λ of the microcapsule was enhanced by 195% when the hybrid filler content reached 0.6 wt.%. This illustrated that GO and the polymer shell exhibited a good interface interaction, whereas CNT presented a poor interface interaction with the shell. To study the dispersion of the fillers in the polymer shell, they poured the prepolymer of the shell with brown GO or black CNTs into a cylinder mold and then cured and dried it, as shown in Figure 20. The results indicate that GO exhibits a better dispersion in the polymer shell than CNT. In addition, the GO–CNT hybrid filler can be dispersed uniformly in the polymer shell because the presence of CNT can promote the dispersion of GO in the shell, leading to homogeneous dispersion of the filler network.

5.2.3. High Thermal Conductivity Shell

In fact, it is difficult to ensure a uniform distribution of heat conduction enhancement particles within the shell, and long-term use will also lead to adverse results such as precipitation and aggregation, thus reducing λ and deteriorating the temperature uniformity of the heat storage system [30]. In addition, this type of enhanced micro/nanocapsule may work in the stacking form, whereas when they are used in the working fluid, the effect of λ enhancement is limited because the filler particles are distributed discontinuously in the shell. In contrast, a continuous shell with a high-λ coating on the PCM core surface is better because it can exhibit heat transfer enhancement in all applications, especially in the working fluid. However, for a continuous shell, there are also factors that influence the λ enhancement of the microcapsules. The λ value of the shell is important for the heat transfer process. Zhang et al. [87] and Yu et al. [150] fabricated n-octadecane microcapsules with inorganic shells; when SiO2 was used as the shell with a tested λ of 1.296 W/(m·K), the λ of the obtained microcapsules was 0.6213 W/(m·K); when CaCO3 was used as the shell with a λ of 2.467 W/(m·K), the λ of the obtained microcapsules was 1.674 W/(m·K). Therefore, a highly thermally conductive shell tends to enhance thermal conductivity. In addition, the λ of the microcapsules was also restrained by the shell thickness. Zhang et al. [87] prepared n-octadecane microcapsules with a silica shell (TEOS as the precursor) via a sol-gel process, and the λ of the microcapsules was determined to be 0.6213 W/(m·K) with an n-octadecane/TEOS mass ratio of 50/50, whereas λ was only tested to be 0.3751 W/(m·K) with an n-octadecane/TEOS mass ratio of 70/30, which indicates that the thin and porous shell material is not conducive to heat transfer.

5.2.4. Mass Ratio of Core/Shell

As discussed in Section 5.1.1, the mass ratio of the core/shell materials affects the thermal storage properties, where a high mass ratio often leads to a large encapsulation ratio. In our previous research [164], the relationship between the encapsulation ratio and thermal conductivity was discussed. Three SA/Ag nanocapsules with a SA content of 0.4 g (S1), 0.8 g (S2), and 1.2 g (S3) were prepared via a chemical reduction method, and the encapsulation ratio was increased from 15.3% to 31.7% to 71.4%. The effective thermal conductivity (λeff) of these three samples was also tested, and the results varied from 6.020 to 0.974 W/m·K. This indicates that the thermal conductivity showed the opposite trend with the variation in R. It is easy to understand that a higher R leads to a lower shell mass ratio. While the shell generally has a relatively higher thermal conductivity than the core PCM, it contributes most to the thermal conductivity of micro/nanocapsules. Therefore, micro/nanocapsules prepared by a high mass ratio of the core/shell show a large encapsulation ratio, which is not beneficial to the thermal conductivity.

5.2.5. Summary of the Influence Factors on Thermal Conductivity

Although both highly thermally conductive particles and shells can improve the thermal conductivity of micro/nanocapsules, their limitations are also inevitable. With an increase in the content of the conductive particles or the high conductive shell thickness, the proportion of core to micro/nanocapsules will be reduced, resulting in a reduction in the thermal storage capacities and encapsulation ratio. The higher the particle (or shell) content, the lower the thermal storage capacity and encapsulation ratio. In addition, for high-density materials such as shells or fillers, such as Ag, the high density will not only lead to a reduction in the thermal storage capacities and encapsulation ratio but may also cause deposition problems when used within the working fluid; therefore, the suspension problem should be discussed. In our previous research [164], we indicated that to maintain a good suspension of Ag-coated micro/nanocapsules, the particle size should be controlled below the theoretical suspension critical diameter (TSCD). Therefore, to enhance thermal conductivity, both methods need to be improved. It is likely that encapsulating PCMs with shell materials of high thermal conductivity and low density is an ideal choice at present, while the search for this material and the development of synthetic technology are the main directions.

5.3. Influence Factors on Thermal Reliability

In the studies at present, most of the preparation of micro/nanocapsules has mentioned the test of thermal reliability, based on which the factors affecting thermal reliability can be summarized.

5.3.1. Compactness of the Shell Material

The compactness of the shell material greatly affects the thermal reliability of the micro/nanocapsules. The latent heat loss after the thermal cycling test relies on the core PCM escaping through the pore structure of the shell material. The more mesopores in the shell structure, the better the core content escapes during the heating/cooling process, which eventually leads to poor thermal reliability. Lan et al. [117] performed shell modification of n-octadecane(n-OD)/SiO2 nanocapsules with poly(hydroxylethyl methacrylate) (PHEMA) and PS with 500 thermal cycling tests. After 500 thermal cycles for n-OD/SiO2 nanocapsules, the melting and solidifying latent heats decreased by 9.0 and 8.7%, respectively, whereas for the n-OD/PS–SiO2 and n-OD/PHEMA–SiO2 nanocapsules, the latent heats decreased much less, indicating that polymer–SiO2 hybrid shell material is more suitable than SiO2 for nanocapsules to repeatedly store/release heat in applications. Liu et al. [174] fabricated uniform n-eicosane/meso-silica nanospheres and nanocapsules using a self-templating strategy (Figure 21). Thermal reliability analysis showed that the n-eicosane/meso-silica nanospheres had poor coincidence in the DSC curves, and the altitudes of crystallization and melting peaks gradually decreased during the repeated phase transition. In contrast, the n-eicosane/meso-silica nanocapsules showed good consistency in phase-change performance because their DSC curves almost overlapped with each other during the repeated phase transition process, and no shift was observed in their melting and crystallization peaks. Besides, the melting and solidifying latent heats of the nanospheres indicated a downward trend with thermal cycles and were reduced by 31.1 and 30.7% after 1000 thermal cycles, respectively. However, the melting and solidifying latent heats showed a stable trend with the thermal cycles and only varied by 0.25 and 0.2 J/g, respectively, after 1000 thermal cycles, which illustrated that silica shell could protect the encapsulated n-eicosane for the nanocapsules. In addition, the freezing and melting temperature fluctuations of the nanocapsules with thermal cycles were 0.19 and 0.36 °C, respectively, which was a narrower range than that of the nanospheres. These results confirmed that the nanocapsules demonstrated better thermal reliability than the nanospheres.

5.3.2. The Size and Shell Thickness

In addition to the intrinsic factors of the shell structure, frequent contact between the micro/nanocapsule particles during extensive heating/cooling cycles can also cause the destruction of the capsule structure and the escape of the core PCM, resulting in the loss of latent heat. The size and shell thickness of the micro/nanocapsules are critical for the destruction of the capsule structure. Sukhorukov et al. [175] studied polyelectrolyte capsules with sizes of 10 nm and 10 mm by applying the same force and found that capsules with a size of 10 nm showed a substantially smaller deformation than that of 10 mm capsules. This study verified that nanocapsules are more stable in structure than macro/microcapsules. Theoretically, smaller-sized capsules have better freedom of contact between particles, whereas larger-sized capsules are more likely to be worn out during the heating/cooling process. For larger-sized capsules, a thick shell may be important to compensate for their structural stability. Melanie et al. [176] conducted a mechanical property test with capsule sizes of 10–50 μm and shell thicknesses of 50–200 nm and found that the stiffness of the capsule increased with the shell thickness. Therefore, a small size and thick shell are favorable factors for promoting the thermal reliability of phase-change capsules. From the above analysis, it appears that capsules with a smaller size would show less latent heat loss after extensive thermal cycling processes. However, when the size is sufficiently small, such as nanocapsules, the effect of size on thermal reliability changes. In our previous study [137], we prepared LA/SiO2 nanocapsules of different sizes, and their thermal reliability was tested over 1200 thermal cycles. The results indicated that the percentage of latent heat reduction after the thermal cycling test decreased significantly with increasing particle size (Figure 22), which was mainly due to the decreased specific surface area of the nanocapsules reducing the volatilization of the encapsulated LA. Therefore, nanocapsules with relatively large sizes are the optimum choices for achieving thermal reliability.

5.3.3. Summary of the Influence Factors on Thermal Reliability

From the above analysis, we can conclude that all these factors can be attributed to strengthening the shell structure and improving the freedom of the particles. However, some of the above influencing factors may weaken the thermal storage capacity, such as shell thickness. As discussed previously, thick shells are beneficial to thermal reliability, but thin shells can improve thermal storage capacity. Therefore, there should be a shell balance when considering thermal storage capacity and thermal reliability. Furthermore, although we summarized some intrinsic factors, there are still various factors, such as the application environment, application temperature, and application form. Further research is necessary to determine its thermal reliability.

6. Application Prospects of Solid–Liquid Organic Phase-Change Micro/Nanocapsules

Various applications of phase-change micro/nanocapsules have been studied. For instance, phase-change micro/nanocapsules used in buildings to save energy, in the food industry and textiles to reduce temperature fluctuations, in the thermal management of electronics to cool the device, in solar systems, and in working fluids to improve heat storage or heat transfer efficiency [29,177]. In particular, for phase-change micro/nanocapsules with good thermal performance, such as large latent heat, high thermal conductivity, and long-term cycling performance, they are crucial to the applications of working fluids for solar systems and thermal management of electronics, and other applications depend on the application form because these thermal performances are key factors affecting the thermal charging/discharging rate.

6.1. Applications in Working Fluids

EPCMs can be dispersed in a fluid to form a multiphase medium that integrates heat storage and heat transfer, which are known as latent functional thermal fluids (LFTF) [178]. Figure 23 shows the microcapsule slurry and an SEM image of the microcapsule particles [179]. For this working fluid, water is commonly used as the carrier fluid, and surfactants are used to help micro/nanocapsules disperse well into the carrier fluid. With the large latent heat and high λ of micro/nanocapsules, the working fluid presents good heat transfer and thermal storage performance in various applications.

6.1.1. Air Conditioning, Ceiling Cooling, and Energy Storage

Wang et al. [179] used a C16H34/amino plastic LFTF combined with evaporative cooling technologies to propose a low-energy air-conditioning strategy applied in a cooling ceiling system. The results indicated that applying the new system would have an energy-saving potential as high as 80% in the northern China climate and 10% in the southern China climate. Wang and Niu [180] conducted a further study and designed a new air-conditioning system under the climatic conditions of Hong Kong by combining a cooled ceiling and a C16H34/UF LFTF storage tank. When the system was installed in an office room, the electricity demand during daytime could be reduced by approximately 33%. Concurrently, by utilizing the high-temperature operation of ceiling panels for sensible space cooling load removal, energy savings and cooling demand shifting could be achieved. The results indicated that the small LFTF storage tank could transfer a portion of the cooling load from daytime to nighttime, and the proposed system is regarded as an energy-saving and economical air-conditioning system. The n-paraffin wax microcapsules (2 μm) with a melting point of 8 °C and latent heats of 75.9 J/g dispersed in water were used for cooling at Narita Airport (Tokyo, Japan) [181]. The LFTF was applied to compensate for the decrease in energy capacity when changing refrigerants for environmental reasons. The storage tank has a height of 24.7 m and a diameter of 7.4 m. Cold energy can be stored at night and released during the day. Compared with the old thermal storage system that used external ice melting, the thermal storage density of the LFTF system was reduced by approximately 60%, but the operating cost was reduced by 32%.

6.1.2. Thermal Management of Electronic Devices

LFTF is preferable for the thermal management of electronic devices because it can store and transfer large amounts of heat at near-constant temperatures [182]. Several studies have reported the use of LFTFs in the thermal management of electronic devices.
Ye et al. [183] studied the thermal management of LED with eicosane/polyurea-polyurethane microcapsules (5–30 μm, 37.0 °C, 180.5 J/g) LFTF as a coolant. The experimental results indicated that a system with LFTF as a coolant has a larger cooling capacity than that with water. In particular, at high ambient temperatures, the effect was more significant when using LFTF instead of water, owing to the phase change of the microcapsules. Li et al. [184] established a model with an LFTF as a coolant to study a battery thermal management system. They observed that the maximum temperature of the battery could be further reduced by using LFTF instead of water as the coolant. In addition, the LFTF flow rate, mass fraction, latent heat of microcapsules, and range of melting point significantly affected the performance. Deng et al. [185] established mathematical models to study the heat transfer performance of an LFTF. When an LFTF is used in microchannels for chip cooling, the heat transfer can be enhanced compared with the single-phase fluid used. Bai et al. [186] studied a thermal management system for batteries with an LFTF and minichannel cooling plates by establishing a three-dimensional thermal model. The results illustrated that the LFTF containing 20 wt.% n-octadecane microcapsules and 80 wt.% water presented better performance than pure water, mineral oil, and glycol solution. Zhang et al. [187] thermodynamically assessed the active cooling/heating methods for batteries with an LFTF cycle. The results indicated that the LFTF cycle method exhibited a higher energy efficiency than the refrigerant circulation method. They also conducted a thermodynamic assessment of active cooling/heating methods under extreme conditions for lithium-ion batteries in electric vehicles. Pakrouh et al. [188] conducted a numerical study on battery thermal management with the application of LFTF, and they observed that the maximum temperature and its difference were reduced by 9.21 and 3.07 K compared with that of pure water, whereas the pressure drop and power consumption increased. Sabbah et al. [189] studied the effect of the LFTF on the heat transfer coefficient of a microchannel radiator by simulation, and using the LFTF as a coolant instead of water, they observed an even distribution of temperature and a significant improvement in performance.
In the present study, the LFTF demonstrated a significant role in the thermal management of electronic devices. In particular, for micro/nanocapsules with good thermal performance, large latent heat could promote the micro/nanocapsules to absorb a large amount of heat from the device and maintain its temperature stability, and high λ can enhance the thermal response by fast heat transfer from the device to the core PCM, which avoids the heat absorption delay caused by the slow heat transfer and good thermal reliability, which can ensure a long service life of the micro/nanocapsules. Therefore, micro/nanocapsules with good thermal performance are ideal candidates for the thermal management of electronic devices.

6.2. Applications in Bulk Form

6.2.1. Green Energy-Saving Building

PCMs excel at increasing the energy storage capacity of the building envelope and regulating the air temperature distribution to improve occupancy comfort.
Buildings account for 40% of global energy consumption and 38% of greenhouse gas emissions. Improving building energy efficiency to mitigate global warming is crucial [190]. PCMs can improve the thermal storage capacity of a building enclosure and adjust the air temperature distribution for living comfort [191,192]. Phase-change micro/nanocapsules are also good candidates for this application because the existence of a shell could prevent the interaction between the core PCM and the building matrix material, and the shell can also prevent the leakage of the liquid core PCM.
Shossig et al. [193] studied the incorporation of paraffin microcapsules into gypsum wallboards. The microcapsules had an average diameter of 8 μm and were homogeneously dispersed between gypsum crystals. Compared to the walls without microcapsules, the indoor temperature fluctuations showed a reduction of approximately 2 °C. Arce et al. [194] integrated phase-change microcapsules into the concrete walls of cubicles and studied the effect of awnings added to the cubicles. They observed that the peak temperatures showed a reduction of approximately 6%, and the comfort time was increased by at least 10% with awnings.
Other researchers have designed and developed a range of ceilings, floors, roofs, and walls incorporating PCM for greater energy savings [177,195], which relies on latent heat, λ, the loading amount of micro/nanocapsules, and the thickness of building decoration materials. Moreover, economic evaluation should consider the applications of PCMs and micro/nanocapsules in buildings [196].

6.2.2. Food Industry

Preserving foods under refrigeration and maintaining low temperatures is essential to avoid/delay microbial, physical, and chemical changes; considering this, phase-change micro/nanocapsules may help prevent temperature fluctuations [197].
Pérez-Masiá et al. [198] produced dodecane/zein micro/nanocapsules using electrospinning. They proposed that the produced structures could act as new smart packaging materials to improve the maintain/control temperature. Unal et al. [199] studied the packaging design with the thermal buffering effect using octanoic acid/styrene polymer microcapsules (ΔHm = 42.9 J/g). The results indicated that using microcapsules or PCMs could provide a thermal buffering effect as long as 8.8 or 6 h for 160 g of chocolate compared with that without PCMs.
In the food industry, PCMs are commonly used in heat storage and transport systems, such as refrigeration, heat treatment sections, and packaging applications, and thus, phase-change micro/nanocapsules show promising prospects for this application.

6.2.3. Temperature-Controlled Textiles

PCMs can be used in temperature-controlled textiles to achieve automatic temperature regulation by storing and releasing thermal energy. Organic PCMs incorporated into textiles provide a comfortable microclimate environment for the human body and thus improve their work efficiency and quality of life [200,201]. In the application of PCM technology in clothing and household products, phase-change microcapsules are combined with acrylic fibers, polyurethane foams, and coating compounds for topical application to fabrics or foams [202,203]. There are various outdoor apparel products with phase-change microcapsules on the market, such as ski suits, hunting suits, boots, gloves, and earmuffs, under the trade names Outlast and Comfor Temp [204].
Shin et al. [204] developed thermoregulating textile materials by adding the prepared eicosane/MF microcapsules to polyester knit fabrics. The strengthening of the microcapsules was sufficient to ensure their stability under stirring in hot water and alkaline solutions. The thermal storage capacity of the temperature-regulating fabric was 0.91–4.44 J/g, which depended on the concentration of microcapsules. After five times of launderings, the treated fabric retained 40% of its heat storage capacity. Paula et al. [205] studied thermo-regulating textiles with paraffin wax microcapsules, and the results indicated that the coating fabric showed a thermal storage capacity of 7.6 J/g with 35% of microcapsules added. They also found that after washing, rub fastness, and ironing treatments, the coating fabric could still maintain high durability and adequate stability. Koo et al. [206] applied phase-change microcapsules to waterproof nylon fabrics using a dual-coating method.
Most organic-based microcapsules can be used to fabricate temperature-controlled textiles; however, it should be noted that their components are environmentally friendly. Some heat-volatile substances, such as formaldehyde, glutaraldehyde, and ether, are harmful to human health [207]. The application of environment-friendly cross-linking agents or shell materials such as chitosan and polyaniline in the preparation of micro/nanocapsules is encouraged [208].

6.2.4. Solar System

Solar energy is regarded as one of the most important sources of renewable energy. Solar water heating (SWH) systems are simple, reliable, and cost-effective for the use of solar energy in hot water production. Energy can be collected and stored so that it can be used when there is no sunlight. SWH systems can be active or passive, depending on the availability of electric pumps in the system. These SWH modes are available in vacuum, flat plates, or centralized collectors. Integrating the PCM in the upper part of the solar water heater tank improves the functionality of the system, as it can enhance the stratification effect. The supply of hot water was from the upper part, where the temperature was the highest (Figure 24). Connecting the inlet of the solar collector to the lower temperature bottom of the tank would produce better collector efficiency. Several factors affect the thermal energy storage system; in particular, the PCM plays an important role in the system.
Su et al. [209] prepared paraffin wax/MF phase-change microcapsules for SWH systems. A theoretical evaluation of the thermal properties of the microcapsules with the highest latent heat (126 J/g) was conducted. Compared with water-based systems, the microcapsule-applied SWH system has a larger TES density and a correspondingly smaller physical storage size. In addition, although the λ of the microcapsules is slightly lower, it is still approximately twice that of the most common storage units with PCM. Furthermore, it compresses the entire system volume using phase-change microcapsules.
In addition, the production of phase-change micro/nanocapsules is beneficial for increasing the efficiency of solar TES by converting solar radiation into thermal energy. Wang et al. [146] reported that n-octadecane/poly(melanine-formaldehyde)-SiC microcapsules (3 μm) have a good storage capacity (167–168 J/g), and they can efficiently perform photothermal conversion and storage of solar energy, especially with 7% of SiC nanoparticles. The λ of the microcapsules was improved by 60.34%, which would result in a temperature of approximately 10 °C higher than that without SiC after irradiating the same time with an infrared lamp.

7. Further Outlook

Phase-change micro/nanocapsules have excellent potential in various fields. The thermal properties of micro/nanocapsules have made significant progress in the past few decades, and some micro/nanocapsules have been applied in industries and households. Although micro/nanocapsules have excellent advantages, a few challenges must be considered.
Although most of the preparation of micro/nanocapsules has mentioned the thermal reliability test, the thermal cycling performance only involved representing the evaluation under a certain number of thermal cycles, whereas the actual service life of the micro/nanocapsules cannot be estimated based on the data. Therefore, finding a unified standard to evaluate the thermal reliability of different phase-change micro/nanocapsules is a problem that needs to be solved.
The preparation technology for micro/nanocapsules is mostly concentrated in the laboratory stage; realizing large-scale production and promoting technology requires further research.
In addition, micro/nanoencapsulation of PCM has several limitations, such as the selection of the correct material and the high cost of the process. Therefore, simplification of the production methods, improvement in stability, and reduction in encapsulation costs are crucial.
In addition, the development of bifunctional/multifunctional phase-change micro/nanocapsules will significantly expand their scope of application beyond the traditional application field.
Finally, a database of PCMs must be established to facilitate the selection of production methods as well as shell materials.

8. Conclusions

Overall, this study reviews the thermal properties and factors of organic solid–liquid phase-change micro/nanocapsules. The following conclusions were drawn from this study:
The thermal properties of solid–liquid PCMs with paraffin-based PCMs, fatty acids, and other PCMs are summarized. In the past years, n-octadecane, n-eicosane, n-hexaoxane, and n-octadecane have been the most commonly used PCMs among paraffins, with SA, PA, and LA being the most commonly studied fatty acid PCMs.
The methods of micro/nanoencapsulation of organic solid–liquid PCMs and their effects on the shell material, size, and properties of the micro/nanocapsules were analyzed. The sol-gel method is used to prepare micro/nanocapsules with organic shells, whereas other methods are mostly used to prepare organic shells. However, all methods can be used to prepare micro/nanocapsules with high thermal storage capacity because they can be adjusted by the preparation conditions. Recently, other synthesis methods have gradually occupied a larger proportion of micro/nanocapsules, indicating that researchers are working on expanding the research on shell materials to further strengthen the performance of micro/nanocapsules.
The thermal properties, including the thermal storage capacity, thermal reliability, and thermal conductivity, of various micro/nanocapsules and their influencing factors were reviewed.
The core/shell mass ratio, shell-forming conditions, and core-formation conditions during the preparation process are the main factors affecting the thermal storage properties. In addition, considering the density difference between the core and shell materials, Rv should be used to fairly assess the encapsulation effect of the micro/nanocapsules.
Nanoparticle content, interface interaction, and dispersion state are key factors affecting the enhanced λ of micro/nanocapsules with additives, whereas a continuous shell with a high λ is better. However, the additive amount of conductive particles is negatively correlated to the thermal storage capacities; this problem requires attention, including the deposition problem when used in the working fluid.
The compactness of the shell, mesopores in the shell structure, and size of the micro/nanocapsule particles are important factors that affect the thermal reliability of micro/nanocapsules.
Finally, the application prospects were presented for phase-change micro/nanocapsules with good thermal performance, which are crucial to the applications of solar systems, working fluids, and thermal management of electronics.

Funding

This study was supported by the National Natural Science Foundation of China (Grant Nos. 52074034 and 52204411).

Data Availability Statement

Not applicable for review paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. da Cunha, J.P.; Eames, P. Thermal energy storage for low and medium temperature applications using phase change materials—A review. Appl. Energy 2016, 177, 227–238. [Google Scholar] [CrossRef]
  2. Du, K.; Calautit, J.; Wang, Z.; Wu, Y.; Liu, H. A review of the applications of phase change materials in cooling, heating and power generation in different temperature ranges. Appl. Energy 2018, 220, 242–273. [Google Scholar] [CrossRef]
  3. Ibrahim, N.I.; Al-Sulaiman, F.A.; Rahman, S.; Yilbas, B.S.; Sahin, A.Z. Heat transfer enhancement of phase change materials for thermal energy storage applications: A critical review. Renew. Sustain. Energy Rev. 2017, 74, 26–50. [Google Scholar] [CrossRef]
  4. Rathod, M.K.; Banerjee, J. Thermal stability of phase change materials used in latent heat energy storage systems: A review. Renew. Sustain. Energy Rev. 2013, 18, 246–258. [Google Scholar] [CrossRef]
  5. Sharma, R.K.; Ganesan, P.; Tyagi, V.V.; Metselaar, H.S.C.; Sandaran, S.C. Developments in organic solid–liquid phase change materials and their applications in thermal energy storage. Energy Convers. Manag. 2015, 95, 193–228. [Google Scholar] [CrossRef]
  6. Shchukina, E.; Graham, M.; Zheng, Z.; Shchukin, D. Nanoencapsulation of phase change materials for advanced thermal energy stor-age systems. Chem. Soc. Rev. 2018, 47, 4156–4175. [Google Scholar] [CrossRef]
  7. Vadhera, J.; Sura, A.; Nandan, G.; Dwivedi, G. Study of Phase Change materials and its domestic application. Mater. Today Proc. 2018, 5, 3411–3417. [Google Scholar] [CrossRef]
  8. Su, W.; Darkwa, J.; Kokogiannakis, G. Review of solid–liquid phase change materials and their encapsulation technologies. Renew. Sustain. Energy Rev. 2015, 48, 373–391. [Google Scholar] [CrossRef]
  9. Giro-Paloma, J.; Martínez, M.; Cabeza, L.F.; Fernández, A.I. Types, methods, techniques, and applications for microencapsulated phase change materials (MPCM): A review. Renew. Sustain. Energy Rev. 2016, 53, 1059–1075. [Google Scholar] [CrossRef]
  10. Meng, J.-Y.; Tang, X.-F.; Zhang, Z.-L.; Zhang, X.-X.; Shi, H.-F. Fabrication and properties of poly (polyethylene glycol octadecyl ether methacrylate). Thermochim. Acta 2013, 574, 116–120. [Google Scholar] [CrossRef]
  11. O’Neil, G.W.; Yen, T.Q.; Leitch, M.A.; Wilson, G.R.; Brown, E.A.; Rider, D.A.; Reddy, C.M. Alkenones as renewable phase change materials. Renew. Energy 2019, 134, 89–94. [Google Scholar] [CrossRef]
  12. Pielichowska, K.; Pielichowski, K. Phase change materials for thermal energy storage. Prog. Mater. Sci. 2014, 65, 67–123. [Google Scholar] [CrossRef]
  13. Yuan, Y.; Zhang, N.; Tao, W.; Cao, X.; He, Y. Fatty acids as phase change materials: A review. Renew. Sustain. Energy Rev. 2014, 29, 482–498. [Google Scholar] [CrossRef]
  14. Ahmed, T.; Bhouri, M.; Groulx, D.; White, M.A. Passive Thermal Management of Tablet PCs Using Phase Change Materials: Intermittent Operation. Appl. Sci. 2019, 9, 902. [Google Scholar] [CrossRef]
  15. Ali, H.M.; Arshad, A.; Janjua, M.M.; Baig, W.; Sajjad, U. Thermal performance of LHSU for electronics under steady and transient oper-ations modes. Int. J. Heat Mass Transfer 2018, 127, 1223–1232. [Google Scholar] [CrossRef]
  16. Setoh, G.; Tan, F.; Fok, S. Experimental studies on the use of a phase change material for cooling mobile phones. Int. Commun. Heat Mass Transf. 2010, 37, 1403–1410. [Google Scholar] [CrossRef]
  17. Tomizawa, Y.; Sasaki, K.; Kuroda, A.; Takeda, R.; Kaito, Y. Experimental and numerical study on phase change material (PCM) for thermal management of mobile devices. Appl. Therm. Eng. 2016, 98, 320–329. [Google Scholar] [CrossRef]
  18. Alva, G.; Lin, Y.; Liu, L.; Fang, G. Synthesis, characterization and applications of microencapsulated phase change materials in ther-mal energy storage: A review. Energy Build. 2017, 144, 276–294. [Google Scholar] [CrossRef]
  19. Chai, L.; Shaukat, R.; Wang, L.; Wang, H.S. A review on heat transfer and hydrodynamic characteristics of nano/microencapsulated phase change slurry (N/MPCS) in mini/microchannel heat sinks. Appl. Therm. Eng. 2018, 135, 334–349. [Google Scholar] [CrossRef]
  20. Chen, L.; Wang, T.; Zhao, Y.; Zhang, X.-R. Characterization of thermal and hydrodynamic properties for microencapsulated phase change slurry (MPCS). Energy Convers. Manag. 2014, 79, 317–333. [Google Scholar] [CrossRef]
  21. Jamekhorshid, A.; Sadrameli, S.M.; Farid, M. A review of microencapsulation methods of phase change materials (PCMs) as a thermal energy storage (TES) medium. Renew. Sustain. Energy Rev. 2014, 31, 531–542. [Google Scholar] [CrossRef]
  22. Qiu, Z.; Ma, X.; Li, P.; Zhao, X.; Wright, A. Micro-encapsulated phase change material (MPCM) slurries: Characterization and build-ing applications. Renew. Sustain. Energy Rev. 2017, 77, 246–262. [Google Scholar] [CrossRef]
  23. Serale, G.; Cascone, Y.; Capozzoli, A.; Fabrizio, E.; Perino, M. Potentialities of a Low Temperature Solar Heating System Based on Slurry Phase Change Materials (PCS). Energy Procedia 2014, 62, 355–363. [Google Scholar] [CrossRef]
  24. Emam, M.; Ahmed, M. Cooling concentrator photovoltaic systems using various configurations of phase-change material heat sinks. Energy Convers. Manag. 2018, 158, 298–314. [Google Scholar] [CrossRef]
  25. Su, Y.; Zhang, Y.; Shu, L. Experimental study of using phase change material cooling in a solar tracking concentrated photovolta-ic-thermal system. Solar Energy 2018, 159, 777–785. [Google Scholar] [CrossRef]
  26. Al Shannaq, R.; Farid, M. Microencapsulation of phase change materials (PCMs) for thermal energy storage systems. In Advances in Thermal Energy Storage Systems; Elsevier: Amsterdam, The Netherlands, 2015; pp. 247–284. [Google Scholar] [CrossRef]
  27. Liu, L.; Alva, G.; Huang, X.; Fang, G. Preparation, heat transfer and flow properties of microencapsulated phase change materials for thermal energy storage. Renew. Sustain. Energy Rev. 2016, 66, 399–414. [Google Scholar] [CrossRef]
  28. Peng, H.; Zhang, D.; Ling, X.; Li, Y.; Wang, Y.; Yu, Q.; She, X.; Li, Y.; Ding, Y. Yulong n-Alkanes Phase Change Materials and Their Microencapsulation for Thermal Energy Storage: A Critical Review. Energy Fuels 2018, 32, 7262–7293. [Google Scholar] [CrossRef]
  29. Leong, K.Y.; Abdul Rahman, M.R.; Gurunathan, B.A. Nano-enhanced phase change materials: A review of thermo-physical properties, applications and challenges. J. Energy Storage 2019, 21, 18–31. [Google Scholar] [CrossRef]
  30. Lin, Y.; Jia, Y.; Alva, G.; Fang, G. Review on thermal conductivity enhancement, thermal properties and applications of phase change materials in thermal energy storage. Renew. Sustain. Energy Rev. 2018, 82, 2730–2742. [Google Scholar] [CrossRef]
  31. Zhang, N.; Yuan, Y.; Cao, X.; Du, Y.; Zhang, Z.; Gui, Y. Latent Heat Thermal Energy Storage Systems with Solid–Liquid Phase Change Materials: A Review. Adv. Eng. Mater. 2018, 20, 1700753. [Google Scholar] [CrossRef]
  32. Qureshi, Z.A.; Ali, H.M.; Khushnood, S. Recent advances on thermal conductivity enhancement of phase change materials for energy storage system: A review. Int. J. Heat Mass Transf. 2018, 127, 838–856. [Google Scholar] [CrossRef]
  33. Yuan, K.; Shi, J.; Aftab, W.; Qin, M.; Zou, R. Engineering the Thermal Conductivity of Functional Phase-Change Materials for Heat Energy Conversion, Storage, and Utilization. Adv. Funct. Mater. 2020, 30, 1904228. [Google Scholar] [CrossRef]
  34. Rodriguez-Cumplido, F.; Pabon-Gelves, E.; Chejne-Jana, F. Recent developments in the synthesis of microencapsulated and nano-encapsulated phase change materials. J. Energy Storage 2019, 24, 100821. [Google Scholar] [CrossRef]
  35. Kenisarin, M.M. Thermophysical properties of some organic phase change materials for latent heat storage. A review. Sol. Energy 2014, 107, 553–575. [Google Scholar] [CrossRef]
  36. Gulfam, R.; Zhang, P.; Meng, Z. Advanced thermal systems driven by paraffin-based phase change materials—A review. Appl. Energy 2019, 238, 582–611. [Google Scholar] [CrossRef]
  37. Singh, J.; Parvate, S.; Vennapusa, J.R.; Maiti, T.K.; Dixit, P.; Chattopadhyay, S. Facile method to prepare 1-dodecanol@ poly (mela-mine-paraformaldehyde) phase change energy storage microcapsules via surfactant-free method. J. Energy Storage 2022, 49, 104089. [Google Scholar] [CrossRef]
  38. Ali, M.A.; Fayaz; Viegas, R.F.; Kumar, M.B.S.; Kannapiran, R.K.; Feroskhan, M. Enhancement of heat transfer in paraffin wax PCM using nano graphene composite for industrial helmets. J. Energy Storage 2019, 26, 100982. [Google Scholar] [CrossRef]
  39. Vakhshouri, A.R. Paraffin as phase change material. Paraffin Overv. 2020, 1–23. [Google Scholar] [CrossRef]
  40. Ohlrogge, J.B.; Jaworski, J.G. Regulation of fatty acid synthesis. Annu. Rev. Plant Biol. 1997, 48, 109–136. [Google Scholar] [CrossRef]
  41. Harwood, J.L. Fatty acid biosynthesis. In Plant Lipids; Blackwell: Tokyo, Japan, 2020; pp. 27–66. [Google Scholar]
  42. Feldman, D.; Shapiro, M.; Banu, D.; Fuks, C. Fatty acids and their mixtures as phase-change materials for thermal energy storage. Sol. Energy Mater. 1989, 18, 201–216. [Google Scholar] [CrossRef]
  43. Hasan, A.; McCormack, S.; Huang, M.; Sarwar, J.; Norton, B. Increased photovoltaic performance through temperature regulation by phase change materials: Materials comparison in different climates. Sol. Energy 2015, 115, 264–276. [Google Scholar] [CrossRef]
  44. Sarı, A.; Karaipekli, A. Fatty acid esters-based composite phase change materials for thermal energy storage in buildings. Appl. Therm. Eng. 2012, 37, 208–216. [Google Scholar] [CrossRef]
  45. Alkan, C.; Sari, A. Fatty acid/poly(methyl methacrylate) (PMMA) blends as form-stable phase change materials for latent heat thermal energy storage. Sol. Energy 2008, 82, 118–124. [Google Scholar] [CrossRef]
  46. Gbabode, G.; Negrier, P.; Mondieig, D.; Moreno Calvo, E.; Calvet, T.; Cuevas-Diarte, M.À. Structures of the High-Temperature Solid Phases of the Odd-Numbered Fatty Acids from Tridecanoic Acid to Tricosanoic Acid. Chemistry 2007, 13, 3150–3159. [Google Scholar] [CrossRef] [PubMed]
  47. Suppes, G.; Goff, M.; Lopes, S. Latent heat characteristics of fatty acid derivatives pursuant phase change material applications. Chem. Eng. Sci. 2003, 58, 1751–1763. [Google Scholar] [CrossRef]
  48. Costa, M.C.; Sardo, M.; Rolemberg, M.P.; Coutinho, J.A.; Meirelles, A.J.; Ribeiro-Claro, P.; Krähenbühl, M.A. The solid–liquid phase diagrams of binary mixtures of consecutive, even saturated fatty acids. Chem. Phys. Lipids 2009, 160, 85–97. [Google Scholar] [CrossRef] [PubMed]
  49. Mishra, R.K.; Verma, K.; Mishra, V.; Chaudhary, B. A review on carbon-based phase change materials for thermal energy storage. J. Energy Storage 2022, 50, 104166. [Google Scholar] [CrossRef]
  50. A Talja, R.; Roos, Y.H. Phase and state transition effects on dielectric, mechanical, and thermal properties of polyols. Thermochim. Acta 2001, 380, 109–121. [Google Scholar] [CrossRef]
  51. Kaizawa, A.; Maruoka, N.; Kawai, A.; Kamano, H.; Jozuka, T.; Senda, T.; Akiyama, T. Thermophysical and heat transfer properties of phase change material candidate for waste heat transportation system. Heat Mass Transf. 2007, 44, 763–769. [Google Scholar] [CrossRef]
  52. Sari, A.; Eroglu, R.; Biçer, A.; Karaipekli, A. Synthesis and Thermal Energy Storage Properties of Erythritol Tetrastearate and Erythritol Tetrapalmitate. Chem. Eng. Technol. 2010, 34, 87–92. [Google Scholar] [CrossRef]
  53. Sari, A.; Karaipekli, A.; Eroğlu, R.; Biçer, A. Erythritol Tetra Myristate and Erythritol Tetra Laurate as Novel Phase Change Materials for Low Temperature Thermal Energy Storage. Energy Sources Part A Recover. Util. Environ. Eff. 2013, 35, 1285–1295. [Google Scholar] [CrossRef]
  54. Nikolić, R.; Marinović-Cincović, M.; Gadžurić, S.; Zsigrai, I. New materials for solar thermal storage—solid/liquid transitions in fatty acid esters. Sol. Energy Mater. Sol. Cells 2003, 79, 285–292. [Google Scholar] [CrossRef]
  55. Chen, C.; Wang, L.; Huang, Y. Electrospun phase change fibers based on polyethylene glycol/cellulose acetate blends. Appl. Energy 2011, 88, 3133–3139. [Google Scholar] [CrossRef]
  56. Constantinescu, M.; Dumitrache, L.; Constantinescu, D.; Anghel, E.M.; Popa, V.T.; Stoica, A.; Olteanu, M. Latent heat nano composite building materials. Eur. Polym. J. 2010, 46, 2247–2254. [Google Scholar] [CrossRef]
  57. Mathur, S.; Moudgil, B.M. Adsorption Mechanism(s) of Poly(Ethylene Oxide) on Oxide Surfaces. J. Colloid Interface Sci. 1997, 196, 92–98. [Google Scholar] [CrossRef]
  58. Meng, Q.; Hu, J. A poly(ethylene glycol)-based smart phase change material. Sol. Energy Mater. Sol. Cells 2008, 92, 1260–1268. [Google Scholar] [CrossRef]
  59. Sarier, N.; Onder, E. Organic phase change materials and their textile applications: An overview. Thermochim. Acta 2012, 540, 7–60. [Google Scholar] [CrossRef]
  60. Sarı, A.; Biçer, A.; Karaipekli, A. Synthesis, characterization, thermal properties of a series of stearic acid esters as novel solid–liquid phase change materials. Mater. Lett. 2009, 63, 1213–1216. [Google Scholar] [CrossRef]
  61. Kou, Y.; Wang, S.; Luo, J.; Sun, K.; Zhang, J.; Tan, Z.; Shi, Q. Thermal analysis and heat capacity study of polyethylene glycol (PEG) phase change materials for thermal energy storage applications. J. Chem. Thermodyn. 2019, 128, 259–274. [Google Scholar] [CrossRef]
  62. Zare, P.; Perera, N.; Lahr, J.; Hasan, R. Solid-liquid phase change materials for the battery thermal management systems in electric vehicles and hybrid electric vehicles—A systematic review. J. Energy Storage 2022, 52, 105026. [Google Scholar] [CrossRef]
  63. Liu, H.; Wang, X.; Wu, D. Innovative design of microencapsulated phase change materials for thermal energy storage and versatile applications: A review. Sustain. Energy Fuels 2019, 3, 1091–1149. [Google Scholar] [CrossRef]
  64. Palanikkumaran, M.; Gupta, K.K.; Agrawal, A.K.; Jassal, M. Highly stable hexamethylolmelamine microcapsules containing n-octadecane prepared by in situ encapsulation. J. Appl. Polym. Sci. 2009, 114, 2997–3002. [Google Scholar] [CrossRef]
  65. Sarier, N.; Onder, E. The manufacture of microencapsulated phase change materials suitable for the design of thermally enhanced fabrics. Thermochim. Acta 2007, 452, 149–160. [Google Scholar] [CrossRef]
  66. 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 Physicochem. Eng. Asp. 2009, 332, 129–138. [Google Scholar] [CrossRef]
  67. Rao, Y.; Lin, G.; Luo, Y.; Chen, S.; Wang, L. Preparation and thermal properties of microencapsulated phase change material for en-hancing fluid flow heat transfer. Heat Transf. Asian Res. 2007, 36, 28–37. [Google Scholar] [CrossRef]
  68. Du, X.; Fang, Y.; Cheng, X.; Du, Z.; Zhou, M.; Wang, H. Fabrication and Characterization of Flame-Retardant Nanoencapsulated n-Octadecane with Melamine–Formaldehyde Shell for Thermal Energy Storage. ACS Sustain. Chem. Eng. 2018, 6, 15541–15549. [Google Scholar] [CrossRef]
  69. Zhao, C.Y.; Zhang, G.H. Review on microencapsulated phase change materials (MEPCMs): Fabrication, characterization and applications. Renew. Sustain. Energy Rev. 2011, 15, 3813–3832. [Google Scholar] [CrossRef]
  70. Peng, H.; Wang, J.; Zhang, X.; Ma, J.; Shen, T.; Li, S.; Dong, B. A review on synthesis, characterization and application of nanoencapsulated phase change materials for thermal energy storage systems. Appl. Therm. Eng. 2020, 185, 116326. [Google Scholar] [CrossRef]
  71. Bryant, Y.G. Melt spun fibers containing microencapsulated phase change material. In ASME International Mechanical Engineer-ing Congress and Exposition; American Society of Mechanical Engineers: Nashville, TN, USA, 1999; pp. 225–234. [Google Scholar]
  72. Shi, J.; Wu, X.; Sun, R.; Ban, B.; Li, J.; Chen, J. Nano-encapsulated phase change materials prepared by one-step interfacial polymeriza-tion for thermal energy storage. Mater. Chem. Phys. 2019, 231, 244–251. [Google Scholar] [CrossRef]
  73. Zou, G.L.; Tan, Z.C.; Lan, X.Z.; Sun, L.X.; Zhang, T. Preparation and characterization of microencapsulated hexadecane used for thermal energy storage. Chin. Chem. Lett. 2004, 15, 729–732. [Google Scholar]
  74. Yuan, H.; Kalfas, G.; Ray, W. Suspension polymerization. J. Macromol. Sci. Part C Polym. Rev. 1991, 31, 215–299. [Google Scholar] [CrossRef]
  75. You, M.; Zhang, X.; Wang, J.; Wang, X. Polyurethane foam containing microencapsulated phase-change materials with styrene–divinybenzene co-polymer shells. J. Mater. Sci. 2009, 44, 3141–3147. [Google Scholar] [CrossRef]
  76. Zhang, K.; Wang, J.; Xu, L.; Xie, H.; Guo, Z. Preparation and thermal characterization of n-octadecane/pentafluorostyrene nanocap-sules for phase-change energy storage. J. Energy Storage 2021, 35, 102327. [Google Scholar] [CrossRef]
  77. Alehosseini, E.; Jafari, S.M. Micro/nano-encapsulated phase change materials (PCMs) as emerging materials for the food industry. Trends Food Sci. Technol. 2019, 91, 116–128. [Google Scholar] [CrossRef]
  78. Arshady, R. Suspension, emulsion, and dispersion polymerization: A methodological survey. Colloid Polym. Sci. 1992, 270, 717–732. [Google Scholar] [CrossRef]
  79. Sarı, A.; Bicer, A.; Alkan, C.; Özcan, A.N. Thermal energy storage characteristics of myristic acid-palmitic eutectic mixtures encapsu-lated in PMMA shell. Sol. Energy Mater. Sol. Cells 2019, 193, 1–6. [Google Scholar] [CrossRef]
  80. Alay, S.; Göde, F.; Alkan, C. Preparation and characterization of poly (methylmethacrylate-coglycidyl methacrylate)/n-hexadecane nanocapsules as a fiber additive for thermal energy storage. Fibers Polym. 2010, 11, 1089–1093. [Google Scholar] [CrossRef]
  81. Liu, C.; Rao, Z.; Zhao, J.; Huo, Y.; Li, Y. Review on nanoencapsulated phase change materials: Preparation, characterization and heat transfer enhancement. Nano Energy 2015, 13, 814–826. [Google Scholar] [CrossRef]
  82. Rezvanpour, M.; Hasanzadeh, M.; Azizi, D.; Rezvanpour, A.; Alizadeh, M. Synthesis and characterization of micro-nanoencapsulated n-eicosane with PMMA shell as novel phase change materials for thermal energy storage. Mater. Chem. Phys. 2018, 215, 299–304. [Google Scholar] [CrossRef]
  83. Fonseca, G.E.; McKenna, T.F.; Dubé, M.A. Miniemulsion vs. conventional emulsion polymerization for pressure-sensitive adhesives production. Chem. Eng. Sci. 2010, 65, 2797–2810. [Google Scholar] [CrossRef]
  84. Yuan, H.; Bai, H.; Zhang, X.; Zhang, J.; Zhang, Z.; Yang, L. Synthesis and characterization of stearic acid/silicon dioxide nanoencapsules for solar energy storage. Sol. Energy 2018, 173, 42–52. [Google Scholar] [CrossRef]
  85. Wang, L.-Y.; Tsai, P.-S.; Yang, Y.-M. Preparation of silica microspheres encapsulating phase-change material by sol-gel method in O/W emulsion. J. Microencapsul. 2006, 23, 3–14. [Google Scholar] [CrossRef]
  86. Fang, G.; Chen, Z.; Li, H. Synthesis and properties of microencapsulated paraffin composites with SiO2 shell as thermal energy storage materials. Chem. Eng. J. 2010, 163, 154–159. [Google Scholar] [CrossRef]
  87. Zhang, H.; Wang, X.; Wu, D. Silica encapsulation of n-octadecane via sol–gel process: A novel microencapsulated phase-change material with enhanced thermal conductivity and performance. J. Colloid Interface Sci. 2010, 343, 246–255. [Google Scholar] [CrossRef]
  88. Latibari, S.T.; Mehrali, M.; Mehrali, M.; Mahlia, T.M.I.; Metselaar, H.S.C. Synthesis, characterization and thermal properties of nanoen-capsulated phase change materials via sol-gel method. Energy 2013, 61, 664–672. [Google Scholar] [CrossRef]
  89. Latibari, S.T.; Mehrali, M.; Mehrali, M.; Afifi, A.B.M.; Mahlia, T.M.I.; Akhiani, A.R.; Metselaar, H.S.C. Facile synthesis and thermal performances of stea-ric acid/titania core/shell nanocapsules by sol–gel method. Energy 2015, 85, 635–644. [Google Scholar] [CrossRef]
  90. Zhang, Y.; Wang, X.; Wu, D. Design and fabrication of dual-functional microcapsules containing phase change material core and zirconium oxide shell with fluorescent characteristics. Sol. Energy Mater. Sol. Cells 2015, 133, 56–68. [Google Scholar] [CrossRef]
  91. Laghari, I.A.; Samykano, M.; Pandey, A.; Said, Z.; Kadirgma, K.; Tyagi, V. Thermal conductivity and Thermal properties enhancement of Paraffin/Titanium Oxide based Nano enhanced Phase change materials for Energy storage. In Proceedings of the 2022 Advances in Science and Engineering Technology International Conferences (ASET), Dubai, United Arab Emirates, 21–24 February 2022; pp. 1–5. [Google Scholar]
  92. Yuan, H.; Bai, H.; Chen, H.; Zhang, Z.; An, H.; Tian, W. Synthesis and properties of high thermal conductivity Ag shell-coated phase change materials. Renew. Energy 2021, 179, 395–405. [Google Scholar] [CrossRef]
  93. Sarı, A.; Alkan, C.; Altıntaş, A. Preparation, characterization and latent heat thermal energy storage properties of mi-cro-nanoencapsulated fatty acids by polystyrene shell. Appl. Therm. Eng. 2014, 73, 1160–1168. [Google Scholar] [CrossRef]
  94. Sarı, A.; Alkan, C.; Biçer, A.; Altuntaş, A.; Bilgin, C. Micro/nanoencapsulated n-nonadecane with poly (methyl methacrylate) shell for thermal energy storage. Energy Convers. Manag. 2014, 86, 614–621. [Google Scholar] [CrossRef]
  95. Sarı, A.; Alkan, C.; Döğüşcü, D.K.; Kızıl, Ç. Micro/nano encapsulated n-tetracosane and n-octadecane eutectic mixture with polystyrene shell for low-temperature latent heat thermal energy storage applications. Sol. Energy 2015, 115, 195–203. [Google Scholar] [CrossRef]
  96. Zhang, J.; Zhao, T.; Chai, Y.; Wang, L. Preparation and Characterization of High Content Paraffin Wax Microcapsules and Mi-cro/Nanocapsules with Poly Methyl Methacrylate Shell by Suspension-Like Polymerization. Chin. J. Chem. 2017, 35, 497–506. [Google Scholar] [CrossRef]
  97. Yu, F.; Chen, Z.-H.; Zeng, X.-R. Preparation, characterization, and thermal properties of microPCMs containing n-dodecanol by us-ing different types of styrene-maleic anhydride as emulsifier. Colloid Polym. Sci. 2009, 287, 549–560. [Google Scholar] [CrossRef]
  98. Fang, G.; Li, H.; Yang, F.; Liu, X.; Wu, S. Preparation and characterization of nano-encapsulated n-tetradecane as phase change mate-rial for thermal energy storage. Chem. Eng. J. 2009, 153, 217–221. [Google Scholar] [CrossRef]
  99. Nan, G.-H.; Wang, J.-P.; Wang, Y.; Wang, H.; Li, W.; Zhang, X.-X. Preparation and properties of nanoencapsulated phase change materials containing polyaniline. Acta Phys. Chim. Sin. 2014, 30, 338–344. [Google Scholar]
  100. Niu, X.; Xu, Q.; Zhang, Y.; Zhang, Y.; Yan, Y.; Liu, T. Fabrication and Properties of Micro-Nanoencapsulated Phase Change Materials for Internally-Cooled Liquid Desiccant Dehumidification. Nanomaterials 2017, 7, 96. [Google Scholar] [CrossRef] [PubMed]
  101. Hu, X.; Huang, Z.; Yu, X.; Li, B. Preparation and Thermal Energy Storage of Carboxymethyl Cellulose-Modified Nanocapsules. BioEnergy Res. 2013, 6, 1135–1141. [Google Scholar] [CrossRef]
  102. Han, S.; Lyu, S.; Wang, S.; Fu, F. High-intensity ultrasound assisted manufacturing of melamine-urea-formaldehyde/paraffin nanocapsules. Colloids Surfaces A Physicochem. Eng. Asp. 2019, 568, 75–83. [Google Scholar] [CrossRef]
  103. Karthikeyan, M.; Ramachandran, T.; Shanmugasundaram, O. Synthesis, characterization, and development of thermally enhanced cotton fabric using nanoencapsulated phase change materials containing paraffin wax. J. Text. Inst. 2014, 105, 1279–1286. [Google Scholar] [CrossRef]
  104. Zhang, X.-J.; Wang, J.; Zhang, X. Preparation of double-shell nanoencapsulated phase change materials by interfacial polymeriza-tion in an emulsion system. New Chem. Mater. 2011, 39, 45–49. [Google Scholar]
  105. Kook, J.; Cho, W.; Koh, W.; Cheong, I.W.; Kim, J.H. Preparation and characterization of octadecane/polyurea nanocapsule-embedded poly(ethylene oxide) nanofibers. J. Appl. Polym. Sci. 2015, 132, 42539. [Google Scholar] [CrossRef]
  106. Barlak, S.; Sara, O.N.; Karaipekli, A.; Yapıcı, S. Thermal Conductivity and Viscosity of Nanofluids Having Nanoencapsulated Phase Change Material. Nanoscale Microscale Thermophys. Eng. 2016, 20, 85–96. [Google Scholar] [CrossRef]
  107. Shi, J.; Wu, X.; Fu, X.; Sun, R. Synthesis and thermal properties of a novel nanoencapsulated phase change material with PMMA and SiO2 as hybrid shell materials. Thermochim. Acta 2015, 617, 90–94. [Google Scholar] [CrossRef]
  108. Pan, L.; Tao, Q.; Zhang, S.; Wang, S.; Zhang, J.; Wang, S.; Wang, Z.; Zhang, Z. Preparation, characterization and thermal properties of mi-cro-encapsulated phase change materials. Sol. Energy Mater. Sol. Cells 2012, 98, 66–70. [Google Scholar] [CrossRef]
  109. Alkan, C.; Sarı, A.; Karaipekli, A.; Uzun, O. Preparation, characterization, and thermal properties of microencapsulated phase change material for thermal energy storage. Sol. Energy Mater. Sol. Cells 2009, 93, 143–147. [Google Scholar] [CrossRef]
  110. Li, H.; Jiang, M.; Li, Q.; Li, D.; Huang, J.; Hu, W.; Dong, L.; Xie, H.; Xiong, C. Facile preparation and thermal performances of hexadecanol/crosslinked poly-styrene core/shell nanocapsules as phase change material. Polym. Compos. 2014, 35, 2154–2158. [Google Scholar] [CrossRef]
  111. Konuklu, Y.; Paksoy, H.O.; Unal, M. Nanoencapsulation of n-alkanes with poly(styrene-co-ethylacrylate) shells for thermal energy storage. Appl. Energy 2015, 150, 335–340. [Google Scholar] [CrossRef]
  112. Alay, S.; Alkan, C.; Göde, F. Synthesis and characterization of poly (methyl methacrylate)/n-hexadecane microcapsules using dif-ferent cross-linkers and their application to some fabrics. Thermochim. Acta 2011, 518, 1–8. [Google Scholar] [CrossRef]
  113. Alkan, C.; Sarı, 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, 687–692. [Google Scholar] [CrossRef]
  114. Giro-Paloma, J.; Konuklu, Y.; Fernández, A. Preparation and exhaustive characterization of paraffin or palmitic acid microcapsules as novel phase change material. Sol. Energy 2015, 112, 300–309. [Google Scholar] [CrossRef]
  115. Li, M.G.; Zhang, Y.; Xu, Y.H.; Zhang, D. Effect of different amounts of surfactant on characteristics of nanoencapsulated phase-change materials. Polym. Bull. 2011, 67, 541–552. [Google Scholar] [CrossRef]
  116. Sirohi, S.; Singh, D.; Nain, R.; Parida, D.; Agrawal, A.K.; Jassal, M. Electrospun composite nanofibres of PVA loaded with nanoencap-sulated n-octadecane. RSC Adv. 2015, 5, 34377–34382. [Google Scholar] [CrossRef]
  117. Zhu, Y.; Qin, Y.; Wei, C.; Liang, S.; Luo, X.; Wang, J.; Zhang, L. Nanoencapsulated phase change materials with polymer-SiO2 hybrid shell materials: Compositions, morphologies, and properties. Energy Convers. Manag. 2018, 164, 83–92. [Google Scholar] [CrossRef]
  118. Du, X.; Wang, S.; Du, Z.; Cheng, X.; Wang, H. Preparation and characterization of flame-retardant nanoencapsulated phase change materials with poly(methylmethacrylate) shells for thermal energy storage. J. Mater. Chem. A 2018, 6, 17519–17529. [Google Scholar] [CrossRef]
  119. Zhang, G.H.; Bon, S.A.; Zhao, C.-Y. Synthesis, characterization and thermal properties of novel nanoencapsulated phase change ma-terials for thermal energy storage. Solar Energy 2012, 86, 1149–1154. [Google Scholar] [CrossRef]
  120. Tumirah, K.; Hussein, M.Z.; Zulkarnain, Z.; Rafeadah, R. Nano-encapsulated organic phase change material based on copolymer nanocomposites for thermal energy storage. Energy 2014, 66, 881–890. [Google Scholar] [CrossRef]
  121. Fang, Y.; Kuang, S.; Gao, X.; Zhang, Z. Preparation and characterization of novel nanoencapsulated phase change materials. Energy Convers. Manag. 2008, 49, 3704–3707. [Google Scholar] [CrossRef]
  122. Fang, Y.; Yu, H.; Wan, W.; Gao, X.; Zhang, Z. Preparation and thermal performance of polystyrene/n-tetradecane composite nanoen-capsulated cold energy storage phase change materials. Energy Convers. Manag. 2013, 76, 430–436. [Google Scholar] [CrossRef]
  123. Fu, W.; Liang, X.; Xie, H.; Wang, S.; Gao, X.; Zhang, Z.; Fang, Y. Thermophysical properties of n-tetradecane@ polystyrene-silica compo-site nanoencapsulated phase change material slurry for cold energy storage. Energy Build. 2017, 136, 26–32. [Google Scholar] [CrossRef]
  124. Fang, Y.; Liu, X.; Liang, X.; Liu, H.; Gao, X.; Zhang, Z. Ultrasonic synthesis and characterization of polystyrene/n-dotriacontane com-posite nanoencapsulated phase change material for thermal energy storage. Appl. Energy 2014, 132, 551–556. [Google Scholar] [CrossRef]
  125. de Cortazar, M.G.; Rodríguez, R. Thermal storage nanocapsules by miniemulsion polymerization. J. Appl. Polym. Sci. 2013, 127, 5059–5064. [Google Scholar] [CrossRef]
  126. Fuensanta, M.; Paiphansiri, U.; Romero-Sánchez, M.D.; Guillem, C.; López-Buendía, M.; Landfester, K. Thermal properties of a novel nanoencapsulated phase change material for thermal energy storage. Thermochim. Acta 2013, 565, 95–101. [Google Scholar] [CrossRef]
  127. Chen, Z.-H.; Yu, F.; Zeng, X.-R.; Zhang, Z.-G. Preparation, characterization and thermal properties of nanocapsules containing phase change material n-dodecanol by miniemulsion polymerization with polymerizable emulsifier. Appl. Energy 2012, 91, 7–12. [Google Scholar] [CrossRef]
  128. Yu, F.; Chen, Z.-H.; Zeng, X.-R.; Gao, X.-N.; Zhang, Z.-G. Poly(methyl methacrylate) copolymer nanocapsules containing phase-change material (n-dodecanol) prepared via miniemulsion polymerization. J. Appl. Polym. Sci. 2015, 132, 42334. [Google Scholar] [CrossRef]
  129. Wang, Y.; Wang, J.; Nan, G.; Wang, H.; Li, W.; Zhang, X. A Novel Method for the Preparation of Narrow-Disperse Nanoencapsulated Phase Change Materials by Phase Inversion Emulsification and Suspension Polymerization. Ind. Eng. Chem. Res. 2015, 54, 9307–9313. [Google Scholar] [CrossRef]
  130. Liu, Y.; Wei, Y.; Kong, W.; Zong, C. Properties of n-octadecane nanocapsules prepared with anionic polyurethane. Polym. Mater. Als Sci. Eng. 2014, 30, 172–176+181. [Google Scholar]
  131. Lu, J.; Yang, L.I.; Han, Y.; Qian, F. Preparation and properties characterization of microencapsulated phase change materials using acrylic resin copolymers/n-dodecanol. Chin. J. Process Eng. 2019, 19, 617–622. [Google Scholar]
  132. Zhu, Y.; Liang, S.; Wang, H.; Zhang, K.; Jia, X.; Tian, C.; Zhou, Y.; Wang, J. Morphological control and thermal properties of nanoencapsulated n-octadecane phase change material with organosilica shell materials. Energy Convers. Manag. 2016, 119, 151–162. [Google Scholar] [CrossRef]
  133. Liang, S.; Zhu, Y.; Wang, H.; Wu, T.; Tian, C.; Wang, J.; Bai, R. Preparation and Characterization of Thermoregulated Rigid Polyurethane Foams Containing Nanoencapsulated Phase Change Materials. Ind. Eng. Chem. Res. 2016, 55, 2721–2730. [Google Scholar] [CrossRef]
  134. Zhu, Y.; Chi, Y.; Liang, S.; Luo, X.; Chen, K.; Tian, C.; Wang, J.; Zhang, L. Novel metal coated nanoencapsulated phase change materials with high thermal conductivity for thermal energy storage. Sol. Energy Mater. Sol. Cells 2018, 176, 212–221. [Google Scholar] [CrossRef]
  135. Zhu, Y.; Liang, S.; Chen, K.; Gao, X.; Chang, P.; Tian, C.; Wang, J.; Huang, Y. Preparation and properties of nanoencapsulated n-octadecane phase change material with organosilica shell for thermal energy storage. Energy Convers. Manag. 2015, 105, 908–917. [Google Scholar] [CrossRef]
  136. Kalaiselvam, S. Bifunctional nanoencapsulated eutectic phase change material core with SiO2/SnO2 nanosphere shell for ther-mal and electrical energy storage. Mater. Des. 2018, 154, 291–301. [Google Scholar]
  137. Yuan, H.; Bai, H.; Lu, X.; Zhang, X.; Zhang, J.; Zhang, Z.; Yang, L. Size controlled lauric acid/silicon dioxide nanocapsules for thermal energy storage. Sol. Energy Mater. Sol. Cells 2019, 191, 243–257. [Google Scholar] [CrossRef]
  138. Yuan, H.; Bai, H.; Lu, X.; Zhao, X.; Zhang, X.; Zhang, J.; Zhang, Z.; Yang, L. Effect of alkaline pH on formation of lauric acid/SiO2 nanocapsules via sol-gel process for solar energy storage. Sol. Energy 2019, 185, 374–386. [Google Scholar] [CrossRef]
  139. Regin, A.F.; Solanki, S.; Saini, J. Heat transfer characteristics of thermal energy storage system using PCM capsules: A review. Renew. Sustain. Energy Rev. 2008, 12, 2438–2458. [Google Scholar] [CrossRef]
  140. Zhang, X.X.; Fan, Y.F.; Tao, X.; Yick, K.L. Fabrication and properties of microcapsules and nanocapsules containing n-octadecane. Mater. Chem. Phys. 2004, 88, 300–307. [Google Scholar] [CrossRef]
  141. Jin, Z.; Wang, Y.; Liu, J.; Yang, Z. Synthesis and properties of paraffin capsules as phase change materials. Polymer 2008, 49, 2903–2910. [Google Scholar] [CrossRef]
  142. 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. Cells 2009, 93, 1366–1376. [Google Scholar] [CrossRef]
  143. Sánchez-Silva, L.; Rodríguez, J.F.; Carmona, M.; Romero, A.; Sánchez, P. Thermal and morphological stability of polystyrene micro-capsules containing phase-change materials. J. Appl. Polym. Sci. 2011, 120, 291–297. [Google Scholar] [CrossRef]
  144. Yang, Y.; Kuang, J.; Wang, H.; Song, G.; Liu, Y.; Tang, G. Enhancement in thermal property of phase change microcapsules with modi-fied silicon nitride for solar energy. Sol. Energy Mater. Sol. Cells 2016, 151, 89–95. [Google Scholar] [CrossRef]
  145. Jiang, X.; Luo, R.; Peng, F.; Fang, Y.; Akiyama, T.; Wang, S. Synthesis, characterization and thermal properties of paraffin microcap-sules modified with nano-Al2O3. Appl. Energy 2015, 137, 731–737. [Google Scholar] [CrossRef]
  146. Wang, X.; Li, C.; Zhao, T. Fabrication and characterization of poly (melamine-formaldehyde)/silicon carbide hybrid microencap-sulated phase change materials with enhanced thermal conductivity and light-heat performance. Sol. Energy Mater. Sol. Cells 2018, 183, 82–91. [Google Scholar] [CrossRef]
  147. Chen, Z.; Wang, J.; Yu, F.; Zhang, Z.; Gao, X. Preparation and properties of graphene oxide-modified poly(melamine-formaldehyde) microcapsules containing phase change material n-dodecanol for thermal energy storage. J. Mater. Chem. A 2015, 3, 11624–11630. [Google Scholar] [CrossRef]
  148. Park, S.; Lee, Y.; Kim, Y.S.; Lee, H.M.; Kim, J.H.; Cheong, I.W.; Koh, W.G. Magnetic nanoparticle-embedded PCM nanocapsules based on par-affin core and polyurea shell. Colloids Surf. A Physicochem. Eng. Asp. 2014, 450, 46–51. [Google Scholar] [CrossRef]
  149. Li, M.; Chen, M.; Wu, Z. Enhancement in thermal property and mechanical property of phase change microcapsule with modified carbon nanotube. Appl. Energy 2014, 127, 166–171. [Google Scholar] [CrossRef]
  150. Yu, S.; Wang, X.; Wu, D. Microencapsulation of n-octadecane phase change material with calcium carbonate shell for enhancement of thermal conductivity and serving durability: Synthesis, microstructure, and performance evaluation. Appl. Energy 2014, 114, 632–643. [Google Scholar] [CrossRef]
  151. Chai, L.; Wang, X.; Wu, D. Development of bifunctional microencapsulated phase change materials with crystalline titanium di-oxide shell for latent-heat storage and photocatalytic effectiveness. Appl. Energy 2015, 138, 661–674. [Google Scholar] [CrossRef]
  152. Yuan, K.; Wang, H.; Liu, J.; Fang, X.; Zhang, Z. Novel slurry containing graphene oxide-grafted microencapsulated phase change material with enhanced thermo-physical properties and photo-thermal performance. Sol. Energy Mater. Sol. Cells 2015, 143, 29–37. [Google Scholar] [CrossRef]
  153. Wang, T.; Wang, S.; Luo, R.; Zhu, C.; Akiyama, T.; Zhang, Z. Microencapsulation of phase change materials with binary cores and cal-cium carbonate shell for thermal energy storage. Appl. Energy 2016, 171, 113–119. [Google Scholar] [CrossRef]
  154. Zhang, Y.; Wang, X.; Wu, D. Microencapsulation of n-dodecane into zirconia shell doped with rare earth: Design and synthesis of bifunctional microcapsules for photoluminescence enhancement and thermal energy storage. Energy 2016, 97, 113–126. [Google Scholar] [CrossRef]
  155. Abhat, A. Low temperature latent heat thermal energy storage: Heat storage materials. Sol. Energy 1983, 30, 313–332. [Google Scholar] [CrossRef]
  156. Salunkhe, P.B.; Devanuri, J.K. Containers for Thermal Energy Storage. In Micro-and Nano-Containers for Smart Applications; Springer: Singapore, 2022; pp. 289–307. [Google Scholar]
  157. Sharma, S.; Sagara, K. Latent Heat Storage Materials and Systems: A Review. Int. J. Green Energy 2005, 2, 1–56. [Google Scholar] [CrossRef]
  158. Ferrer, G.; Solé, A.; Barreneche, C.; Martorell, I.; Cabeza, L.F. Review on the methodology used in thermal stability characterization of phase change materials. Renew. Sustain. Energy Rev. 2015, 50, 665–685. [Google Scholar] [CrossRef]
  159. Lan, W.; Shang, B.; Wu, R.; Yu, X.; Hu, R.; Luo, X. Thermally-enhanced nanoencapsulated phase change materials for latent function-ally thermal fluid. Int. J. Therm. Sci. 2021, 159, 106619. [Google Scholar] [CrossRef]
  160. Nikpourian, H.; Bahramian, A.R.; Abdollahi, M. On the thermal performance of a novel PCM nanocapsule: The effect of core/shell. Renew. Energy 2020, 151, 322–331. [Google Scholar] [CrossRef]
  161. Dhivya, S.; Hussain, S.I.; Kalaiselvam, S. Novel metal coated nanocapsules of ethyl esters fatty acid eutectic mixture as phase change material with enhanced thermal conductivity for energy storage applications. Thermochim. Acta 2020, 687, 178581. [Google Scholar] [CrossRef]
  162. Wang, H.; Ma, W.; Zhang, J.; Yang, Z.; Zong, D. Novel synthesis of silica coated palmitic acid nanocapsules for thermal energy storage. J. Energy Storage 2020, 30, 101402. [Google Scholar] [CrossRef]
  163. He, L.; Mo, S.; Lin, P.; Jia, L.; Chen, Y.; Cheng, Z. D-mannitol@silica/graphene oxide nanoencapsulated phase change material with high phase change properties and thermal reliability. Appl. Energy 2020, 268, 115020. [Google Scholar] [CrossRef]
  164. Yuan, H.; Liu, S.; Hao, S.; Zhang, Z.; An, H.; Tian, W.; Chan, M.; Bai, H. Fabrication and properties of nanoencapsulated stearic acid phase change materials with Ag shell for solar energy storage. Sol. Energy Mater. Sol. Cells 2022, 239, 111653. [Google Scholar] [CrossRef]
  165. Trivedi, G.; Parameshwaran, R. Micro/nanoencapsulation of dimethyl adipate with melamine formaldehyde shell as phase change material slurries for cool thermal energy storage. Chem. Thermodyn. Therm. Anal. 2022, 6, 100037. [Google Scholar] [CrossRef]
  166. Wang, J.; Zhai, X.; Zhong, Z.; Zhang, X.; Peng, H. Nanoencapsulated n-tetradecane phase change materials with melamine–urea–formaldehyde–TiO2 hybrid shell for cold energy storage. Colloids Surf. A Physicochem. Eng. Asp. 2022, 636, 128162. [Google Scholar] [CrossRef]
  167. Liang, S.; Li, Q.; Zhu, Y.; Chen, K.; Tian, C.; Wang, J.; Bai, R. Nanoencapsulation of n-octadecane phase change material with silica shell through interfacial hydrolysis and polycondensation in miniemulsion. Energy 2015, 93, 1684–1692. [Google Scholar] [CrossRef]
  168. Cárdenas-Ramírez, C.; Gómez, M.A.; Jaramillo, F.; Fernández, A.G.; Cabeza, L.F. Thermal reliability of organic-organic phase change materials and their shape-stabilized composites. J. Energy Storage 2021, 40, 102661. [Google Scholar] [CrossRef]
  169. Hu, D.; Wang, Z.; Ma, W. Fabrication and characterization of a novel polyurethane microencapsulated phase change material for thermal energy storage. Prog. Org. Coatings 2021, 151, 106006. [Google Scholar] [CrossRef]
  170. Liu, C.; Cao, H.; Jin, S.; Bao, Y.; Cheng, Q.; Rao, Z. Synthesis and characterization of microencapsulated phase change material with phenol-formaldehyde resin shell for thermal energy storage. Sol. Energy Mater. Sol. Cells 2022, 243, 111789. [Google Scholar] [CrossRef]
  171. Zeng, J.L.; Cao, Z.; Yang, D.W.; Xu, F.; Sun, L.X.; Zhang, X.F.; Zhang, L. Effects of MWNTs on phase change enthalpy and thermal conductivity of a solid-liquid organic PCM. J. Therm. Anal. Calorim. 2009, 95, 507–512. [Google Scholar] [CrossRef]
  172. Zhang, Q.; Luo, Z.; Guo, Q.; Wu, G. Preparation and thermal properties of short carbon fibers/erythritol phase change materials. Energy Convers. Manag. 2017, 136, 220–228. [Google Scholar] [CrossRef]
  173. Liu, Z.; Chen, Z.; Yu, F. Enhanced thermal conductivity of microencapsulated phase change materials based on graphene oxide and carbon nanotube hybrid filler. Sol. Energy Mater. Sol. Cells 2019, 192, 72–80. [Google Scholar] [CrossRef]
  174. Liu, H.; Niu, J.; Wang, X.; Wu, D. Design and construction of mesoporous silica/n-eicosane phase-change nanocomposites for su-percooling depression and heat transfer enhancement. Energy 2019, 188, 116075. [Google Scholar] [CrossRef]
  175. Sukhorukov, G.; Fery, A.; Möhwald, H. Intelligent micro- and nanocapsules. Prog. Polym. Sci. 2005, 30, 885–897. [Google Scholar] [CrossRef]
  176. Pretzl, M.; Neubauer, M.; Tekaat, M.; Kunert, C.; Kuttner, C.; Leon, G.; Berthier, D.; Erni, P.; Ouali, L.; Fery, A. Formation and Mechanical Characterization of Aminoplast Core/Shell Microcapsules. ACS Appl. Mater. Interfaces 2012, 4, 2940–2948. [Google Scholar] [CrossRef]
  177. Alehosseini, E.; Jafari, S.M. Nanoencapsulation of phase change materials (PCMs) and their applications in various fields for en-ergy storage and management. Adv. Colloid Interface Sci. 2020, 283, 102226. [Google Scholar] [CrossRef]
  178. Zhang, P.; Ma, Z.; Wang, R. An overview of phase change material slurries: MPCS and CHS. Renew. Sustain. Energy Rev. 2010, 14, 598–614. [Google Scholar] [CrossRef]
  179. Wang, X.; Niu, J.; van Paassen, A. Raising evaporative cooling potentials using combined cooled ceiling and MPCM slurry storage. Energy Build. 2008, 40, 1691–1698. [Google Scholar] [CrossRef]
  180. Wang, X.; Niu, J. Performance of cooled-ceiling operating with MPCM slurry. Energy Convers. Manag. 2009, 50, 583–591. [Google Scholar] [CrossRef]
  181. Zhang, P.; Ma, Z. An overview of fundamental studies and applications of phase change material slurries to secondary loop re-frigeration and air conditioning systems. Renew. Sustain. Energy Rev. 2012, 16, 5021–5058. [Google Scholar] [CrossRef]
  182. Pathak, L.; Trivedi, G.; Parameshwaran, R.; Deshmukh, S.S. Microencapsulated phase change materials as slurries for thermal ener-gy storage: A review. Mater. Today Proc. 2021, 44, 1960–1963. [Google Scholar] [CrossRef]
  183. Ye, J.; Mo, S.; Jia, L.; Chen, Y. Experimental performance of a LED thermal management system with suspended microencapsulated phase change material. Appl. Therm. Eng. 2022, 207, 118155. [Google Scholar] [CrossRef]
  184. Li, H.; Xiao, X.; Wang, Y.; Lian, C.; Li, Q.; Wang, Z. Performance investigation of a battery thermal management system with microen-capsulated phase change material suspension. Appl. Therm. Eng. 2020, 180, 115795. [Google Scholar] [CrossRef]
  185. Deng, X.; Wang, S.; Wang, J.; Zhang, T. Analytical modeling of microchannel heat sinks using microencapsulated phase change ma-terial slurry for chip cooling. Procedia Eng. 2017, 205, 2704–27011. [Google Scholar] [CrossRef]
  186. Bai, F.; Chen, M.; Song, W.; Yu, Q.; Li, Y.; Feng, Z.; Ding, Y. Investigation of thermal management for lithium-ion pouch battery module based on phase change slurry and mini channel cooling plate. Energy 2019, 167, 561–574. [Google Scholar] [CrossRef]
  187. Zhang, X.; Kong, X.; Li, G.; Li, J. Thermodynamic assessment of active cooling/heating methods for lithium-ion batteries of electric vehicles in extreme conditions. Energy 2014, 64, 1092–1101. [Google Scholar] [CrossRef]
  188. Pakrouh, R.; Hosseini, M.; Bahrampoury, R.; Ranjbar, A.; Borhani, S. Cylindrical battery thermal management based on microencapsulated phase change slurry. J. Energy Storage 2021, 40, 102602. [Google Scholar] [CrossRef]
  189. Sabbah, R.; Farid, M.M.; Al-Hallaj, S. Micro-channel heat sink with slurry of water with micro-encapsulated phase change material: 3D-numerical study. Appl. Therm. Eng. 2009, 29, 445–454. [Google Scholar] [CrossRef]
  190. Sivanathan, A.; Dou, Q.; Wang, Y.; Li, Y.; Corker, J.; Zhou, Y.; Fan, M. Phase change materials for building construction: An overview of nano-/micro-encapsulation. Nanotechnol. Rev. 2020, 9, 896–921. [Google Scholar] [CrossRef]
  191. Aguayo, M.; Das, S.; Maroli, A.; Kabay, N.; Mertens, J.C.; Rajan, S.D.; Sant, G.; Chawla, N.; Neithalath, N. The influence of microencapsulated phase change material (PCM) characteristics on the microstructure and strength of cementitious composites: Experiments and finite element simulations. Cem. Concr. Compos. 2016, 73, 29–41. [Google Scholar] [CrossRef]
  192. Arıcı, M.; Bilgin, F.; Nižetić, S.; Karabay, H. PCM integrated to external building walls: An optimization study on maximum activation of latent heat. Appl. Therm. Eng. 2020, 165, 114560. [Google Scholar] [CrossRef]
  193. Schossig, P.; Henning, H.-M.; Gschwander, S.; Haussmann, T. Micro-encapsulated phase-change materials integrated into construction materials. Sol. Energy Mater. Sol. Cells 2005, 89, 297–306. [Google Scholar] [CrossRef]
  194. Arce, P.; Castellón, C.; Castell, A.; Cabeza, L.F. Use of microencapsulated PCM in buildings and the effect of adding awnings. Energy Build. 2012, 44, 88–93. [Google Scholar] [CrossRef]
  195. Song, M.; Niu, F.; Mao, N.; Hu, Y.; Deng, S. Review on building energy performance improvement using phase change materials. Energy Build. 2018, 158, 776–793. [Google Scholar] [CrossRef]
  196. Arivazhagan, R.; Geetha, N.; Sivasamy, P.; Kumaran, P.; Gnanamithra, M.K.; Sankar, S.; Loganathan, G.B.; Arivarasan, A. Review on performance assessment of phase change materials in buildings for thermal management through passive approach. Mater. Today Proc. 2020, 22, 419–431. [Google Scholar] [CrossRef]
  197. Singh, S.; Gaikwad, K.K.; Lee, Y.S. Phase change materials for advanced cooling packaging. Environ. Chem. Lett. 2018, 16, 845–859. [Google Scholar] [CrossRef]
  198. Pérez-Masiá, R.; López-Rubio, A.; Lagarón, J.M. Development of zein-based heat-management structures for smart food packaging. Food Hydrocoll. 2013, 30, 182–191. [Google Scholar] [CrossRef]
  199. Ünal, M.; Konuklu, Y.; Paksoy, H. Thermal buffering effect of a packaging design with microencapsulated phase change material. Int. J. Energy Res. 2019, 43, 4495–4505. [Google Scholar] [CrossRef]
  200. Geng, X.; Li, W.; Wang, Y.; Lu, J.; Wang, J.; Wang, N.; Li, J.; Zhang, X. Reversible thermochromic microencapsulated phase change materials for thermal energy storage application in thermal protective clothing. Appl. Energy 2018, 217, 281–294. [Google Scholar] [CrossRef]
  201. Iqbal, K.; Sun, D. Development of thermo-regulating polypropylene fibre containing microencapsulated phase change materials. Renew. Energy 2014, 71, 473–479. [Google Scholar] [CrossRef]
  202. Pause, B. Nonwoven Protective Garments with Thermo-Regulating Properties. J. Ind. Text. 2003, 33, 93–99. [Google Scholar] [CrossRef]
  203. Shin, Y.; Yoo, D.; Jang, J. Molecular weight effect on antimicrobial activity of chitosan treated cotton fabrics. J. Appl. Polym. Sci. 2001, 80, 2495–2501. [Google Scholar] [CrossRef]
  204. Shin, Y.; Yoo, D.; Son, K. Development of thermoregulating textile materials with microencapsulated phase change materials (PCM). II. Preparation and application of PCM microcapsules. J. Appl. Polym. Sci. 2005, 96, 2005–2010. [Google Scholar] [CrossRef]
  205. Sánchez, P.; Sánchez-Fernandez, M.V.; Romero, A.; Rodríguez, J.F.; Sánchez-Silva, L. Development of thermo-regulating textiles using paraffin wax microcapsules. Thermochim. Acta 2010, 498, 16–21. [Google Scholar] [CrossRef]
  206. Koo, K.; Park, Y.; Choe, J.; Kim, E. The application of microencapsulated phase-change materials to nylon fabric using direct dual coating method. J. Appl. Polym. Sci. 2008, 108, 2337–2344. [Google Scholar] [CrossRef]
  207. Tao, J.; Luan, J.; Liu, Y.; Qu, D.; Yan, Z.; Ke, X. Technology development and application prospects of organic-based phase change materials: An overview. Renew. Sustain. Energy Rev. 2022, 159, 112175. [Google Scholar] [CrossRef]
  208. Silakhori, M.; Metselaar, H.; Mahlia, T.; Fauzi, H. Preparation and characterisation of microencapsulated paraffin wax with polyani-line-based polymer shells for thermal energy storage. Mater. Res. Innov. 2014, 18, S6-480–S6-484. [Google Scholar] [CrossRef]
  209. Su, W.; Darkwa, J.; Kokogiannakis, G. Development of microencapsulated phase change material for solar thermal energy storage. Appl. Therm. Eng. 2017, 112, 1205–1212. [Google Scholar] [CrossRef]
Figure 1. Studies related to solid–liquid PCMs within recent years.
Figure 1. Studies related to solid–liquid PCMs within recent years.
Energies 17 00604 g001
Figure 2. Schematic formation of the microcapsules based on an n-octadecane core and resorcinol-modified melamine–formaldehyde shell through in situ polymerization. Reprinted with permission from Ref. [66]. Copyright (2009) Elsevier.
Figure 2. Schematic formation of the microcapsules based on an n-octadecane core and resorcinol-modified melamine–formaldehyde shell through in situ polymerization. Reprinted with permission from Ref. [66]. Copyright (2009) Elsevier.
Energies 17 00604 g002
Figure 3. The schematic formation mechanism of paraffin/PMMA micro/nanocapsules by the interfacial polymerization method. Reprinted with permission from Ref. [72]. Copyright (2019) Elsevier.
Figure 3. The schematic formation mechanism of paraffin/PMMA micro/nanocapsules by the interfacial polymerization method. Reprinted with permission from Ref. [72]. Copyright (2019) Elsevier.
Energies 17 00604 g003
Figure 4. The schematic formation mechanism of phase-change microcapsules by the suspension polymerization method. Reprinted with permission from Ref. [77]. Copyright (2019) Elsevier.
Figure 4. The schematic formation mechanism of phase-change microcapsules by the suspension polymerization method. Reprinted with permission from Ref. [77]. Copyright (2019) Elsevier.
Energies 17 00604 g004
Figure 5. Schematic representation of early stages of emulsion polymerization. Reprinted with permission from Ref. [77]. Copyright (2019) Elsevier.
Figure 5. Schematic representation of early stages of emulsion polymerization. Reprinted with permission from Ref. [77]. Copyright (2019) Elsevier.
Energies 17 00604 g005
Figure 6. Schematic representation of miniemulsion polymerization. Reprinted with permission from Ref. [77]. Copyright (2019) Elsevier.
Figure 6. Schematic representation of miniemulsion polymerization. Reprinted with permission from Ref. [77]. Copyright (2019) Elsevier.
Energies 17 00604 g006
Figure 7. Schematic formation of the SA/SiO2 nanocapsules by the sol-gel process: (a) hydrolysis of TEOS with ammonia used as the catalyst, (b) condensation of the hydrolysis product, and (c) silica clusters coated on the SA emulsion droplets to form the nanocapsules. Reprinted with permission from Ref. [84]. Copyright (2018) Elsevier.
Figure 7. Schematic formation of the SA/SiO2 nanocapsules by the sol-gel process: (a) hydrolysis of TEOS with ammonia used as the catalyst, (b) condensation of the hydrolysis product, and (c) silica clusters coated on the SA emulsion droplets to form the nanocapsules. Reprinted with permission from Ref. [84]. Copyright (2018) Elsevier.
Energies 17 00604 g007
Figure 8. The size (a), encapsulation (b,c) analysis of micro/nanocapsules with different methods, as well as the preparation trend in the recent five years (d,e). Research percentage indicates the percentage of studies in each year compared to the total number of studies surveyed.
Figure 8. The size (a), encapsulation (b,c) analysis of micro/nanocapsules with different methods, as well as the preparation trend in the recent five years (d,e). Research percentage indicates the percentage of studies in each year compared to the total number of studies surveyed.
Energies 17 00604 g008
Figure 9. The micro/nanocapsules with morphology and DSC results: (a) fatty acid/PS microcapsules, (b) n-nonadecane/PMMA microcapsules, and (c) C24–C18 eutectic mixture/PS microcapsules. Reprinted with permission from Refs. [93,94,95]. Copyright (2014, 2015) Elsevier.
Figure 9. The micro/nanocapsules with morphology and DSC results: (a) fatty acid/PS microcapsules, (b) n-nonadecane/PMMA microcapsules, and (c) C24–C18 eutectic mixture/PS microcapsules. Reprinted with permission from Refs. [93,94,95]. Copyright (2014, 2015) Elsevier.
Energies 17 00604 g009
Figure 10. Schematic of synthetic strategy and the morphology for microcapsules with modified silicon nitride. Reprinted with permission from Ref. [144]. Copyright (2016) Elsevier.
Figure 10. Schematic of synthetic strategy and the morphology for microcapsules with modified silicon nitride. Reprinted with permission from Ref. [144]. Copyright (2016) Elsevier.
Energies 17 00604 g010
Figure 11. The procedures and chemical reactions for PDA modification and silver coating of the NanoPCMs and the SEM with TEM images of the corresponding NanoPCMs prepared with different AgNO3 concentrations. Reprinted with permission from Ref. [134]. Copyright (2018) Elsevier.
Figure 11. The procedures and chemical reactions for PDA modification and silver coating of the NanoPCMs and the SEM with TEM images of the corresponding NanoPCMs prepared with different AgNO3 concentrations. Reprinted with permission from Ref. [134]. Copyright (2018) Elsevier.
Energies 17 00604 g011
Figure 12. Schematic diagram of the synthesis process and the morphology of Ag shell-coated phase-change micro/nanocapsules. Reprinted with permission from Ref. [92]. Copyright (2021) Elsevier.
Figure 12. Schematic diagram of the synthesis process and the morphology of Ag shell-coated phase-change micro/nanocapsules. Reprinted with permission from Ref. [92]. Copyright (2021) Elsevier.
Energies 17 00604 g012
Figure 13. DSC curves of melting and crystallization of paraffin wax/PU nanocapsules under uncycled, 50th, and 100th thermal cycles. Reprinted with permission from Ref. [160]. Copyright (2020) Elsevier.
Figure 13. DSC curves of melting and crystallization of paraffin wax/PU nanocapsules under uncycled, 50th, and 100th thermal cycles. Reprinted with permission from Ref. [160]. Copyright (2020) Elsevier.
Energies 17 00604 g013
Figure 14. DSC curves for the melting (a) and solidifying (b) processes of pure methyl laurate and microcapsules synthesized at different core/shell mass ratios of 1:1, 2:1, 3:1, and 4:1. Reprinted with permission from Ref. [169]. Copyright (2021) Elsevier.
Figure 14. DSC curves for the melting (a) and solidifying (b) processes of pure methyl laurate and microcapsules synthesized at different core/shell mass ratios of 1:1, 2:1, 3:1, and 4:1. Reprinted with permission from Ref. [169]. Copyright (2021) Elsevier.
Energies 17 00604 g014
Figure 15. DSC curves for the melting (a) and solidifying (b) processes of pure SA and the SA/TiO2 nanocapsules with different SA/TTIP ratios: S1 (50/50), S2 (60/40), S3 (70/30), and S4 (80/20). Reprinted with permission from Ref. [89]. Copyright (2015) Elsevier.
Figure 15. DSC curves for the melting (a) and solidifying (b) processes of pure SA and the SA/TiO2 nanocapsules with different SA/TTIP ratios: S1 (50/50), S2 (60/40), S3 (70/30), and S4 (80/20). Reprinted with permission from Ref. [89]. Copyright (2015) Elsevier.
Energies 17 00604 g015
Figure 16. Scheme of the fabrication route for n-eicosane/PF microcapsules. Reprinted with permission from Ref. [170]. Copyright (2022) Elsevier.
Figure 16. Scheme of the fabrication route for n-eicosane/PF microcapsules. Reprinted with permission from Ref. [170]. Copyright (2022) Elsevier.
Energies 17 00604 g016
Figure 17. Analysis of the relationship between particle size and latent heats of the LA/SiO2 nanocapsules. Reprinted with permission from Ref. [137]. Copyright (2019) Elsevier.
Figure 17. Analysis of the relationship between particle size and latent heats of the LA/SiO2 nanocapsules. Reprinted with permission from Ref. [137]. Copyright (2019) Elsevier.
Energies 17 00604 g017
Figure 18. Analysis of the LA/SiO2 nanocapsules: (a) the size distribution of LA emulsion droplets, (b) the relation of the core diameter with the encapsulation ratio, and (c) the relation of shell thickness with the core diameter. Reprinted with permission from Ref. [137]. Copyright (2019) Elsevier.
Figure 18. Analysis of the LA/SiO2 nanocapsules: (a) the size distribution of LA emulsion droplets, (b) the relation of the core diameter with the encapsulation ratio, and (c) the relation of shell thickness with the core diameter. Reprinted with permission from Ref. [137]. Copyright (2019) Elsevier.
Energies 17 00604 g018
Figure 19. Comparison of R and Rv with different k: (a) structural demonstration for R and Rv, and (b) detailed relation analysis with graphs.
Figure 19. Comparison of R and Rv with different k: (a) structural demonstration for R and Rv, and (b) detailed relation analysis with graphs.
Energies 17 00604 g019
Figure 20. SEM images of (a) GO and (b) CNT and fracture morphology of (c) MF, (d) MF/GO, (e) MF/CNT, and (fh) MF/GO-CNT. Reprinted with permission from Ref. [173]. Copyright (2019) Elsevier.
Figure 20. SEM images of (a) GO and (b) CNT and fracture morphology of (c) MF, (d) MF/GO, (e) MF/CNT, and (fh) MF/GO-CNT. Reprinted with permission from Ref. [173]. Copyright (2019) Elsevier.
Energies 17 00604 g020
Figure 21. Formation mechanisms schematic of the meso-silica/n-eicosane nanospheres and nanocapsules. Reprinted with permission from Ref. [174]. Copyright (2019) Elsevier.
Figure 21. Formation mechanisms schematic of the meso-silica/n-eicosane nanospheres and nanocapsules. Reprinted with permission from Ref. [174]. Copyright (2019) Elsevier.
Energies 17 00604 g021
Figure 22. Damping decrement for nanocapsules with different mean diameters. Reprinted with permission from Ref. [137]. Copyright (2019) Elsevier.
Figure 22. Damping decrement for nanocapsules with different mean diameters. Reprinted with permission from Ref. [137]. Copyright (2019) Elsevier.
Energies 17 00604 g022
Figure 23. Appearance of the microcapsule’s slurry and an SEM microscopic image of the MPCM particles. Reprinted with permission from Ref. [179]. Copyright (2008) Elsevier.
Figure 23. Appearance of the microcapsule’s slurry and an SEM microscopic image of the MPCM particles. Reprinted with permission from Ref. [179]. Copyright (2008) Elsevier.
Energies 17 00604 g023
Figure 24. The schematic of a solar hot water system. Reprinted with permission from Ref. [177]. Copyright (2020) Elsevier.
Figure 24. The schematic of a solar hot water system. Reprinted with permission from Ref. [177]. Copyright (2020) Elsevier.
Energies 17 00604 g024
Table 1. Physical properties of some paraffin waxes [39].
Table 1. Physical properties of some paraffin waxes [39].
ParaffinsTm (°C)ΔHm (J/g)λ (W/m·K)
n-Tetradecane (C14)6228–2300.14
n-Pentadecane (C15)102050.2
n-Hexadecane (C16)182370.2
n-Heptadecane (C17)222130.145
n-Octadecane (C18)282450.148
n-Nonadecane (C19)322220.22
n-Eicosane (C20)37246
n-Henicosane (C21)40200, 213
n-Docosane (C22)44.52490.2
n-Tricosane (C23)47.5232
n-Tetracosane (C24)52255
n-Pentacosane (C25)54238
n-Hexacosane (C26)56.5256
n-Heptacosane (C27)59236
n-Octacosane (C28)64.5253
n-Nonacosane (C29)65240
n-Triacontane (C30)66251
n-Hentriacontane (C31)67242
n-Dotriacontane (C32)69170
n-Triatriacontane (C33)712680.2
Paraffin C16-C1820–22152
Paraffin C13-C2422–241890.21
RT 35 HC352400.2
Paraffin C16-C2842–44189
Paraffin C20-C3348–50189
Paraffin C22-C4558–601890.2
Paraffin C21-C5066–68189
RT 70 HC69–712600.2
Paraffin natural wax 81182–86850.72
Table 2. Thermal properties of some fatty acids [5,48,49].
Table 2. Thermal properties of some fatty acids [5,48,49].
Fatty AcidTm (°C)ΔHm (J/g)λ (W/m·K)
Enanthic−7.4107
Butyric−5.6126
Caproic−3131
Propyl palmiate10186
Pelargonic12.3127
Isopropyl stearate14–18140–142
Phenylacetic16.7102
Caprylic161490.149
Butyl stearate19140
Dimethyl sabacate21120~135
Undecylenic24.6141
Vinyl stearate27~29122
Undecylic28.4139
Capric31.51530.153
Tridecylic41.8157
Methyl-12 hydroxy-stearate42~43120~126
Lauric451930.150
Elaidic47218
Myristic562120.150
Pentadecanoic52~53178
Margaric60172
Palmitic acid622180.162
Stearic acid712370.172
Nonadecylic67192
Arachidic74227
Heneicosylic73~74193
Acetamide81241
Table 3. Thermal properties of sugar alcohol, ester, and glycols.
Table 3. Thermal properties of sugar alcohol, ester, and glycols.
Tm (°C)ΔHm (J/g)λ (W/m·K)Reference
Erythritol118.4379.6 [52,53]
Butyl stearate 23.71210.230[60]
Isopropyl stearate 22.1113.10.150
Glycerol tristearate 63.5149.40.170
Erythritol tetrapalmitate21.9 201.1 0.250[44]
Erythritol tetrastearate 30.4 208.8 0.260
PEG4003.291.4 [59]
PEG100034.8154.40.280[61]
PEG150047.2161.40.310
PEG200050.8165.40.310
PEG400056.0173.60.330
PEG600059.5179.70.340
PEG800059.7177.50.330
PEG1000058.0182.90.330
PEG1200060.9173.40.320
PEG2000062.3168.50.320
Table 4. Volumetric energy storage density of the phase-change micro/nanocapsules.
Table 4. Volumetric energy storage density of the phase-change micro/nanocapsules.
EPCMsR (%)ΔHm, PCMs (J/g) ρ (g/cm3)ΔHv (J/cm3)Reference
CA/PS45.4086.450.9783.86[93]
LA/PS43.5687.210.9683.72
MA/PS48.7098.260.9492.36
C19/PMMA60.30139.200.91126.67[94]
PS/(C24-C18)(2:1)29.6071.730.9668.86[95]
PS/(C24-C18)(1:1)48.0116.320.91105.95
PS/(C24-C18)(1:2)64.4156.390.87136.06
Paraffin wax/PMMA (94 μm)89.5137.200.92126.22[96]
Paraffin wax/PMMA (0.1–19 μm)80.2123.000.94115.62
n-dodecanol/MF93.1187.500.85159.38[97]
C12/MF60.0150.000.87130.50[67]
C14/UF61.8134.160.88118.06[98]
Table 5. Research on the phase-change micro/nanocapsules with thermal storage properties.
Table 5. Research on the phase-change micro/nanocapsules with thermal storage properties.
ShellCoreMeltingSolidifyingR (%)D (nm)Reference
Tm (°C)ΔHm (J/g)Tc (°C)ΔHc (J/g)
UFC149.01134.162.81134.5061.80100.00[98]
8.49131.092.19129.8360.30
7.19103.571.78103.4847.70
5.5766.012.4066.3930.40
PNDA (modified MF resin)C1832.60126.1018.90124.3087.00109.60[68]
32.50125.4019.40123.5086.40109.30
32.00126.5019.70123.7086.90108.80
32.40124.5019.50123.4086.10110.20
32.10125.9019.70124.1086.80111.70
P (MMA-co-AMA)C1824.70129.0023.80155.0058.11692.80[99]
24.60140.0025.3165.0063.06
25.20140.0025.7156.0063.06614.00
27.40151.0027.2166.0068.02
26.50141.0026.6161.0063.51577.50
MF resinC1834.70195.0032.10182.0087.00840.00[99]
34.4097.0015.50101.0045.00
MF resinC1828.06137.6225.82137.9359.58484.00[100]
28.14145.2625.32147.0162.88636.00
28.22137.0725.26137.8659.34630.00
27.88117.1225.51115.5150.70684.00
27.62111.3625.15109.7848.21723.00
PNDA (modified MF resin)Paraffin28.0083.46NANA62.0050.00[101]
MUFParaffin27.5049.4030.1046.0022.80413.30± 1.30[102]
27.4069.2029.9067.5031.60340.40 ± 3.20
27.1080.3030.2075.1036.70370.40 ± 6.40
27.3062.0029.9056.9028.30368.00 ± 2.30
28.3024.9029.7024.5014.90455.70 ± 25.10
UFParaffin64.3074.20NANA48.12256.00[103]
PU and PSC1828.00122.0022.00125.0054.00410.00[104]
PUC1823.70123.4028.20124.1078.10200.00[105]
PUC1930.5492.8530.7891.8644.70103.00[106]
PMMA
SiO2
Paraffin26.8069.9019.8071.0057.40120.00[107]
26.6057.5019.6057.5046.90
27.1053.5019.5053.5043.60
28.0062.1019.8062.4050.70
AlOOHPA12.7019.00NANA10.19200.00[108]
13.8022.60NANA12.12
16.0027.80NANA14.91
PMMAC2241.0054.6040.6048.7028.00160.00[109]
P (MMA-co-GMA)C1617.23148.0514.85147.6362.46260.00[80]
PSHexadecanol52.68124.85NANAc48.72120.00[110]
PS-co-EAC147.97182.68−0.18184.9479.2650–200[111]
C1511.60121.835.20127.8969.16
C1620.38196.0912.78201.7087.09
C1724.04140.5117.23149.6081.59
PMMAC1617.34145.6114.85128.1961.42220[112]
PMMAC2035.2084.2034.9087.5035700[113]
PS-co-EAParaffin42.3949.0337.4149.0532.12165[114]
PA62.6697.9356.2297.2147.79265
PMMAC1931.23139.2031.03142.3960.30 [94]
PSC24-C1825.96156.3926.04152.8364.40 [95]
UFC1616.15143.7016.04134.3068.90253.00[115]
16.26121.0016.32112.5058.06259.00
16.27119.4015.88112.3057.29268.00
16.36114.6016.0798.9954.99285.00
PSC1827.0044.0023.0043.0020.4731.00[116]
SiO2C1827.50108.605.6099.8053.80335.00[117]
PHEMA-SiO227.90104.308.60104.8051.60449.00
PS-SiO227.90101.107.0097.3050.10289.00
PMMAC1832.60110.1019.50108.3091.00104.00[118]
33.50111.2018.30110.0092.20101.00
32.70108.9018.80107.2090.10109.00
33.20110.5019.00108.6091.30112.00
32.00109.1018.80107.2090.10106.00
PEMAC1832.20198.5029.80197.1089.50140[119]
PMMA31.90208.730.2205.994.7119
St-MMAC1830.2097.9025.9093.8040.90152.00 ± 12.00[120]
28.9074.4025.6073.8031.10236.00 ± 13.00
28.4085.9025.3082.9035.90146.00 ± 51.00
29.50107.9024.60104.9045.10102.00 ± 11.00
30.00117.3026.70113.8049.00127.00 ± 30.00
PSC1820–25124.40NANA53.55100–123[121]
PSC144.0498.71−3.4391.2789.00132.00[122]
PS-SiO2C142.1383.380.3979.3742.56120.00[123]
PS2.5982.350.00881.4642.0480.00
PMMAC2034.66124.7032.92119.1362.00135.00[82]
PSC3270.90174.8061.80177.161.23168.00[124]
St-co-MMAParaffin61.50140.30NANA60.70439.40[125]
60.4079.75NANA34.50274.80
60.4044.21NANA19.10118.80
61.4017.20NANA7.5074.00
Styrene–butylRT8078.404.9054.106.503.0452.00[126]
acrylate81.8016.6055.0018.9010.358.00
copolymer79.8012.0054.4014.607.4562.00
PMMAn-dodecanol18.2098.80NANA82.20150.00[127]
PMMAn-dodecanol18.40109.3NANA44.0750–100[128]
MMAC1833.20116.0015.70112.0052.90466.00[129]
PUTC1831.3035.2019.5034.8055.00601.00[130]
30.1067.7020.0065.4074.00698.00
30.8081.8019.9080.1070.00723.00
28.4036.4018.1035.4070.00432.00
29.5048.2019.3047.7062.00598.00
30.1073.7020.3072.8071.00654.00
25.3018.5015.4018.2072.00401.00
29.0039.6018.3038.6076.00432.00
31.9058.5017.9057.5075.00497.00
Acrylic resin copolymersn-dodecanol22.2693.31NANA43.00933.91[131]
OrganosilicaC1828.44113.3023.76108.8055.40NA[132]
SiO2C1827.35116.0024.17109.00NA563.00[133]
Ag-coated SiO2C1828.51126.9022.19120.6062.83274.00[134]
28.27106.5822.13104.8052.77
28.0371.9522.3071.6535.62
28.0549.3022.0350.6024.41
27.7733.1922.2134.0016.43
27.5126.1322.2626.8712.93
OrganosilicaC1827.92107.5024.58102.0051.26385.00[135]
27.9298.4124.9492.3246.93346.00
27.69105.5025.12102.4050.30624.00
27.3593.2024.1691.7344.44693.00
26.5195.1823.3192.9845.39200.00
SiO2PEG3.1962.272.6157.5355.20528.00[136]
SiO2PA61.60180.9157.08181.2289.55722.50[88]
60.92172.1656.80173.4085.22466.40
61.06168.1657.62170.2383.25183.70
SiO2LA36.70165.6034.30152.5085.90210–460[137]
SiO2LA34.5040.6029.4032.5021.1060.00[138]
37.3090.4032.8082.3046.90104.00
36.80116.0034.8094.0060.20221.00
38.80160.0032.90154.8083.00357.00
36.70120.5034.30119.4062.50615.00
38.807.2030.204.703.70750.00
40.604.3028.802.102.20828.00
TiO2SA59.14123.9650.54109.4364.76946.40[89]
58.7299.0250.2084.9351.73620.10
58.4687.9549.9774.7645.94583.40
58.2358.1249.8050.5930.36317.60
NA: no data.
Table 6. Summary of micro/nanocapsules for thermal conductivity enhancement.
Table 6. Summary of micro/nanocapsules for thermal conductivity enhancement.
CoreλPCM (W/(m·K))ShellDR (%)λmicro/nanocapsules
(W/(m·K))
Enhancement
(%)
Reference
n-octadecane PMF 77.00About 0.1300 [146]
PMF + nano-SiC (3 wt%)72.90 18.31
PMF + nano-SiC (5 wt%)72.80 41.36
PMF + nano-SiC (7 wt%)72.30 60.34
n-dodecanolPMF 0.11570[147]
PMF+GO (0.2 wt%) 0.129712.1
PMF+GO (0.4 wt%) 0.139620.65
PMF+GO (0.6 wt%) 0.158637.07
PMF+GO (0.8 wt%) 0.166343.73
PMF+GO (1 wt%) 0.173850.21
PMF+GO (2 wt%) 0.189763.95
PMF+GO (3 wt%) 0.191865.77
PMF+GO (4 wt%) 0.192466.29
Paraffin0.220Polyurea400~600 nm66.560.2250102[148]
Polyurea + nano-Fe3O4 (3.1 wt%)62.110.2330106
Polyurea + nano-Fe3O4 (5.7 wt%)56.420.3350152
Polyurea + nano-Fe3O4 (6.6 wt%)54.830.3420155
Paraffin0.220PMMA0.5~2 μm63.590.2442111[145]
PMMA + nano-Al2O3 (5 wt%)60.770.2786127
PMMA + nano-Al2O3 (17 wt%)53.810.3104141
PMMA + nano-Al2O3 (27 wt%)48.700.3409155
PMMA + nano-Al2O3 (33 wt%)43.920.3591163
PMMA + nano-Al2O3 (38 wt%)43.430.3816173
n-octadecane0.150PMMA4~11 μm73.14 [144]
PMMA + Si3N4 (5 wt%)88.000.2315154
PMMA + Si3N4 (10 wt%)81.920.2753184
PMMA + Si3N4 (15 wt%)73.780.2918195
PMMA + Si3N4 (20 wt%)66.130.2997200
PMMA + Si3N4 (30 wt%)63.610.3630242
Paraffin UFR13.08 μm37.380.0820 [149]
UFR + CNT (1 wt%)0.15 μm23.140.0950
UFR + CNT (2 wt%) 0.1170
UFR + CNT (3 wt%) 0.1330
UFR + CNT (4 wt%) 0.1460
n-octadecane0.150SiO27~16 μm 0.6213414[87]
n-octadecane0.150CaCO35 μm21.891.67401120[150]
n-eicosane0.161TiO21.5~2 μm49.900.8650540[151]
n-octadecane0.153SiO2274 nm62.830.2460161[134]
SiO2 + PDA300 nm52.770.1980129
SiO2 + PDA + Ag (8 g/L)500 nm35.620.5360350
SiO2 + PDA + Ag (16 g/L)550 nm24.410.5180339
SiO2 + PDA + Ag (32 g/L)650 nm16.430.8400549
SiO2 + PDA + Ag (48 g/L)600 nm12.941.3460161
Paraffin0.300SiO230 μm50.801.0310344[152]
SiO2 + GO 49.601.1620387
RT280.289CaCO3 0.7590262[153]
RT28+RT42 CaCO3 0.7390
RT420.388CaCO3 0.9360241
n-docecane0.152ZrO21~1.5 μm64.40.906596[154]
PA0.226Ag100~600 nm34.304.07201800[92]
Table 7. Thermal reliability data with the related references about solid–liquid organic micro/nanocapsules.
Table 7. Thermal reliability data with the related references about solid–liquid organic micro/nanocapsules.
CoreShellDTm (°C)ΔHm (J/g)Tc (°C)ΔHc (J/g)NReference
ES-EPSiO2-Ag510~580 nm20.2108.515.0107.80[161]
20.1108.414.9107.7200
20.1108.414.9107.6500
20.0108.314.8107.61000
SiO2-Cu510~580 nm20.6109.715.6108.80
20.5109.715.6108.8200
20.5109.615.5108.7500
20.4109.615.5108.71000
LASiO2131 nm35.936.833.748.61200[137]
142 nm37.081.933.176.1
174 nm37.2107.533.089.3
311 nm36.9165.136.1150.5
408 nm35.8133.433.4117.4
510 nm37.0115.233.5112.3
474 nm35.1130.733.2127.4
250 nm36.8120.334.5112.7
402 nm37.9143.533.3120.6
215 nm36.5115.533.0103.8
122 nm37.527.531.125.5
n-octadecanePPFS490 nm31.2171.823.0169.3200[76]
PASiO2474 nm62.9109.957.8100.10[162]
62.8109.257.998.6100
PAAg100–600 nm56.40107.8757.46105.222000[92]
D-mannitolsilica/GO100–400 nm166.2216.7122.3174.41[163]
205.6 176.710
214.6 181.120
196.0 166.630
199.5 169.940
206.7 169.250
Paraffin waxPU25~185 nm62.4153.966.9142.30[160]
63.0152.266.8142.450
63.0151.266.8141.5100
SAAg167~252 nm46.13117.7366.13116.612000[164]
SATiO2946 nm59.14123.9550.54109.430[89]
59.41119.2350.01105.682500
PASiO2221 nm38.8160.032.8154.80[138]
38.9157.332.9152.53000
357 nm36.7120.534.3119.40
36.8118.434.3116.73000
615 nm36.8116.034.894.00
36.9114.034.792.83000
Dimethyl adipateMF900 nm6.480.2 100[165]
C14MUF209 nm7.7140.3−0.76142.2100[166]
MUF-TiO2156.2 nm8.4156.2−4.32156.0100
n-octadecaneSiO2346 nm27.9298.424.9492.3500[135]
99.1 92.3100
101.6 94.4300
101.6 94.9500
n-octadecaneSiO2624 nm27.69105.525.12102.40[135]
107.6 104.9100
105.1 102.9300
102.7 100.7500
n-octadecaneST-MMA102 nm24.9107.927.3104.90[120]
24.8106.727.3105.990
24.4105.127.5100.5180
24.9105.427.8101.4270
24.1105.127.1103.7360
n-nonadecanePMMA8.18 μm31.2139.231.0142.40[94]
31.2132.7 5000
CAPS7.7 μm21.8980.56 0[93]
17.2280.28 5000
LAPS13.2 μm21.8980.56 0
17.2280.28 5000
MAPS42.0 μm47.1292.12 0
48.7488.29 5000
n-octadecaneSiO2563 nm27.35109.524.1798.850[167]
109.4 99.091
118.2 107.3100
115.3 103.4300
114.4 102.4500
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yuan, H.; Liu, S.; Li, T.; Yang, L.; Li, D.; Bai, H.; Wang, X. Review on Thermal Properties with Influence Factors of Solid–Liquid Organic Phase-Change Micro/Nanocapsules. Energies 2024, 17, 604. https://doi.org/10.3390/en17030604

AMA Style

Yuan H, Liu S, Li T, Yang L, Li D, Bai H, Wang X. Review on Thermal Properties with Influence Factors of Solid–Liquid Organic Phase-Change Micro/Nanocapsules. Energies. 2024; 17(3):604. https://doi.org/10.3390/en17030604

Chicago/Turabian Style

Yuan, Huanmei, Sitong Liu, Tonghe Li, Liyun Yang, Dehong Li, Hao Bai, and Xiaodong Wang. 2024. "Review on Thermal Properties with Influence Factors of Solid–Liquid Organic Phase-Change Micro/Nanocapsules" Energies 17, no. 3: 604. https://doi.org/10.3390/en17030604

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop