Next Article in Journal
Using Plantain Rachis Fibers and Mopa-Mopa Resin to Develop a Fully Biobased Composite Material
Previous Article in Journal
Controlled Alignment of Carbon Black Nanoparticles in Electrospun Carbon Nanofibers for Flexible Films
Previous Article in Special Issue
Thermal Performance of Multifunctional Facade Solution Containing Phase Change Materials: Experimental and Numerical Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bio-Based Polymers for Environmentally Friendly Phase Change Materials

by
Kinga Pielichowska
1,*,
Katarzyna Nowicka-Dunal
1 and
Krzysztof Pielichowski
2
1
Faculty of Materials Science and Ceramics, Department of Biomaterials and Composites, AGH University of Krakow, al. Mickiewicza 30, 30-059 Kraków, Poland
2
Faculty of Chemical Engineering and Technology, Department of Chemistry and Technology of Polymers, Cracow University of Technology, ul. Warszawska 24, 31-155 Kraków, Poland
*
Author to whom correspondence should be addressed.
Polymers 2024, 16(3), 328; https://doi.org/10.3390/polym16030328
Submission received: 1 December 2023 / Revised: 14 January 2024 / Accepted: 20 January 2024 / Published: 25 January 2024
(This article belongs to the Special Issue Polymers and Phase Change Materials: Past, Present and Future)

Abstract

:
Phase change materials (PCMs) have received increasing attention in recent years as they enable the storage of thermal energy in the form of sensible and latent heat, and they are used in advanced technical solutions for the conservation of sustainable and waste energy. Importantly, most of the currently applied PCMs are produced from non-renewable sources and their carbon footprint is associated with some environmental impact. However, novel PCMs can also be designed and fabricated using green materials without or with a slight impact on the environment. In this work, the current state of knowledge on the bio-based polymers in PCM applications is described. Bio-based polymers can be applied as phase-change materials, as well as for PCMs encapsulation and shape stabilization, such as cellulose and its derivatives, chitosan, lignin, gelatin, and starch. Vast attention has been paid to evaluation of properties of the final PCMs and their application potential in various sectors. Novel strategies for improving their thermal energy storage characteristics, as well as to impart multifunctional features, have been presented. It is also discussed how bio-based polymers can extend in future the potential of new environmentally-safe PCMs in various industrial fields.

1. Introduction

A crucial problem of humanity within the 21st century is fossil fuels’ exhaustion and global warming that forces people to search for new solutions for energy saving and the production of energy from renewable sources.
Phase-change materials (PCMs) are an important class of thermoresponsive materials used for the storage of thermal energy as sensible and latent heat. The application of PCMs in energy-related technical solutions can substantially impact the efficient use and conservation of sustainable and waste energy. Thermal energy storage in the form of latent heat provides a large density of energy storage with small temperature fluctuations during heat storage and release. As PCMs, different types of materials have been investigated including salts, salt hydrates, paraffins, and fatty acids, as well as polymeric materials usually characterized by solid–solid or solid–liquid phase transition [1,2,3]. They are usually modified with different materials to obtain shape-stabilized PCMs, encapsulated PCMs, or to improve some properties, e.g., thermal conductivity.
Most of the currently applied PCMs are obtained from non-renewable sources and their environmental impact is connected with a certain carbon footprint. However, novel PCMs can also be designed and fabricated using green materials without or with a slight impact on the environment.
Polymers are used in PCMs in different aspects, e.g., as PCMs themselves, for PCMs’ shape stabilization, or microencapsulation, but also as precursors in the carbonization process to obtain porous carbon materials that can be then applied for the shape stabilization of PCMs. Representative examples are poly(ethylene glycol) that is widely used as a polymeric PCM, and phenol formaldehyde and melamine formaldehyde resins, polyester, and acrylate resins that are widely applied for the microencapsulation of paraffin or fatty-acid-based PCMs [4]. On the other hand, for the shape stabilization of PCMs polyurethanes and polyethylene are utilized [5].
To decrease the negative impact of PCMs on the natural environment and carbon footprint, there is a search for “green’ solutions. One of the potential routes is the application of bio-based polymeric materials in the broad areas of PCMs and thermal energy storage by replacing (at least partially) synthetic polymers with bio-based ones [6,7,8]. Following this route it is possible to minimize PCMs environmental impact and carbon footprint in various sectors, such as building and energy sectors.
Hence, this work reviews the current state of knowledge on bio-based polymers in PCM applications. Special attention has been paid to evaluation of properties of the final PCMs, and their application potential in various sectors. Novel strategies for improving their TES characteristics, as well as to impart multifunctional features, have been presented. The strategies for improving the thermal energy storage characteristics include attempts to (i) impart multifunctional features and (ii) to minimize PCMs environmental impact and carbon footprint. By using bio-based polymers it is possible to minimize negative environmental impact of PCMs. For this purpose bio-based polymers can be applied both as PCMs themselves and for PCMs shape stabilization. On the other hand, some additional but important features can be added by PCM modification with other additives to fabricate composite materials devoted for e.g., photothermal energy conversion, thermoelectric materials or materials for biomedical applications. It is also discussed how bio-based polymers can extend the potential of new environmentally safe PCMs in various industrial fields.

2. Bio-Based Polymers

According to the definition, bio-based polymers are polymers obtained from renewable resources, mostly from biomass [9]. At the end of the 20th century, the polymer industry was confronted with global warming, as well as the depletion of crude oil and petroleum-based products. Nowadays, bio-based polymers are a subject of intensive research, as they can play a crucial role in decreasing the carbon dioxide emissions causing global warming and using limited fossil resources [10]. One of the possible solutions of these problems was to use sustainable resources and materials instead of components derived from fossil resources [11,12]. The natural source of raw materials is biomass, and agricultural waste from, e.g., corn or potatoes could be a feedstock for numerous bio-based polymers [13].
However, in recent years, numerous research and industrial efforts have been made to develop new biotechnology and green chemical processes in order to preserve land devoted for food production.
Bio-based polymers can be divided in three main groups:
  • First class—polymers prepared from biomass, such as starch, cellulose, chitosan, chitin, sodium alginate, and natural rubber, including those polymers which are chemically modified;
  • Second class—polymers obtained using microorganisms and plants, e.g., poly(hydroxyalkanoates) and poly(glutamic acid);
  • Third class—synthetic polymers obtained from naturally derived monomers from renewable resources, e.g., bio-polyolefins, polylactide, poly(lactide-co-glycolide), poly(butylene succinate), bio-poly(ethylene glycol) PEG, bio-poly(ethylene terephthalate) [9,11,14].
Moreover, in agreement with the circular economy concept, the entire polymeric materials value chain, including materials’ design, recovery, and recycling after their lifetime, should be considered (Figure 1).
Polymer materials based on renewable feedstocks are already being considered as carbon-neutral alternatives to fossil-based polymers, and renewable energy production plays a more and more important role in modern electricity systems.
Bio-based polymers can be biodegradable (polyhydroxyalkanoates, polylactide, starch, alginates) or non-biodegradable (bio-polyolefins, bio-based PEG). The classification of biopolymers is presented in Figure 2.
The biopolymers market has considerably grown recently up to the global production of about 2.4 million tons expected in the current year. However, it is still less than one percent of the total polymer production p.a. (390 million tons in 2021 [17]). Starch and starch-based polymers, polylactide (PLA), and polybutylene adipate terephthalate (PBAT) are the most commonly produced bio-based biodegradable polymers, while the group of bio-based non-biodegradable polymers includes PE, polyamide (PA), poly(ethylene terephthalate) (PET), and poly(trimethylene terephthalate) (PTT). Non-biodegradable biopolymers account for 44.5% of the annual production of biopolymers, while biodegradable biopolymers account for 55.5% of global production [16]. However, for PCMs applications, biodegradability is important only for some applications, e.g., in the biomedical field. For technical and industrial applications, usually non-biodegradable bio-based polymers are required. Some examples of the bio-based polymers currently available on the market are presented in Table 1; in Table 2, PCM systems with bio-based polymer components are shown.

3. Bio-Based Polymers as Phase-Change Materials

Different types of polymers able to achieve a high degree of crystallinity can be applied as PCMs. In this section, polymers that are commercially available and produced using synthetic routes, which can also be obtained as bio-based polymers, have been described. In polymers, the key issue determining the temperature and heat of phase transition is the degree of crystallinity, which suggests that promising candidates for PCM will be polymers composed of linear macrochains, as well as those in which intermolecular forces enhance the arrangement of crystalline structures. In these polymers which could also be renewably sourced, crystallization occurs relatively easily and leads to the formation of regular crystalline forms. Hence, Kamimioto et al. [66] investigated the possibility of using high-density polyethylene (HDPE) as a PCM. HDPE melts at a temperature of 125–135 °C with a melting heat of up to 240 J/g. An additional advantage of using HDPE as a PCM is the possibility of crosslinking it; and then at the melting temperature of linear HDPE, a solid–solid transition of the crosslinked polymer is observed, associated with changes in macromolecular arrangement. Moreover, HDPE crosslinking (using physical or chemical methods) makes it possible to maintain shape stability at temperatures above the melting point.
In another work, Weingrill et al. [67] investigated high-density polyethylene (HDPE) long-term behavior (up to 7200 h) under air atmosphere. They found a minor loss in the storage capacity in thermo-oxidative conditions and concluded that HDPE is a good candidate material for polymeric PCM. Following this research program, the same research team analyzed the applicability of different semi-crystalline polymers, characterized by different phase-transition temperature ranges, as phase-change materials. As promising polymers for PCM applications, they proposed HDPE, PP, POM-H, and POM-C, as well as PA 4.6 and PA 6 [68]. The examined semi-crystalline polymers are presented in Figure 3.
Another polymer considered as PCM was trans-1,4-polybutadiene (PB), which, depending on the temperature, occurs in two crystalline forms [69]. The solid–solid phase transition was detected at 78.1 °C, and the heat of this transformation was 144 J/g. Notably, PB melting occurs at a much higher temperature of 140 °C.
A promising group of macromolecular compounds to be applied as PCMs are those based on polyethers, especially polyethylene oxide (PEO). Aliphatic polyethers, also known as poly(alkylene oxides) or poly(alkylene glycols) (when the macromolecules are terminated with an –OH group) or polyalkyl oxides, are a group of polymers with a general formula –O–R–, where R consists of at least two methylene groups, formed by the ring-opening ionic polymerization of oxiranes and using initiating compounds containing an active hydrogen atom [70]. The most important representatives of aliphatic polyethers are PEO, poly(propylene oxide) (PPO), and polytetrahydrofurane (PTHF) [71,72].
Polyethers’ properties depend on the structure and average molecular weight. The first in the homologous series of aliphatic polyethers–poly(ethylene oxide) with an average molar mass not exceeding 600 is an odorless liquid with high viscosity at room temperature, while with a molar mass higher than 1000 it is a solid. The melting point ranges from 1–5 °C for PEO 400 to 65–67 °C for PEO 100,000—Table 3 [73].
Poly(trimethylene oxide) (PTMO) with three methylene units between ether oxygen atoms has a melting point ca. 35 °C, while polytetrahydrofurane with four methylene units between ether oxygen atoms shows a melting temperature ca. 60 °C. In aliphatic polyethers with an increasing number of methylene groups between ether oxygen atoms, the melting point rises to the value of ca. 130 °C, characteristic for linear polyethylene. All unsubstituted poly(alkylene oxides) are highly crystalline, with a degree of crystallinity, even up to 95% for PEO, but the presence of a side chain in the polyether chain causes in extreme cases a completely amorphous structure. PEO is soluble in water and polar organic solvents, and the chemical properties of polyethers are determined by the presence of ether bonds and hydroxyl end groups. Hydroxyl end groups undergo all reactions characteristic of alcohols, while ether bonds are highly resistant to hydrolysis [71,74]. The thermal properties of selected polyethers are shown in Table 4.

4. PCMs Encapsulation Using Bio-Based Polymers

In most applications, PCMs with a solid–liquid phase transition cannot be directly used due to the leakage above their melting point. Among the most common and effective methods to avoid the leakage, as well as to enhance other properties (such as mechanical ones), is encapsulation; in this process, PCM is sealed by selected tough materials that additionally allow for the enhancement of some PCM properties. The encapsulated PCMs usually consist of a polymer shell and core (the PCM itself). As a shell material, one can also use organic, inorganic, or hybrid organic–inorganic materials. The shell material selection should take into account its compatibility with PCM but can also aim at improving selected properties, e.g., dimensional stability or thermal conductivity [76,77]. The advantages and disadvantages of encapsulation protocols for PCMs are summarized in Table 5.
Among synthetic polymer-based shell materials, melamine-formaldehyde (MF) resin, poly(urea-urethane) (PUU), polyurea (PU), urea-formaldehyde (UF) resin, acrylic, resins and polystyrene are vastly used. However, numerous bio-based polymers have been used as shell materials for PCMs, including gum Arabic, chitosan/gelatin, and agar/gelatin/gum Arabic compositions [77].
Hence, Deveci and Basal [78] prepared microcapsules with an n-eicosane core as a PCM and silk fibroin (SF) and chitosan (CS) as a shell material by the complex coacervation method. Sphere-shaped microcapsules were successfully obtained, and the EDS analysis confirmed the effective incorporation of n-eicosane in the microcapsules. The microcapsules obtained at an SF/CS ratio of 5 possessed a dense and nonporous wall, while for the ratio of 20 an inner layer with a smoother surface and a rough outer surface with a spongelike structure were observed. It has also been found that the microencapsulation efficiency affected the n-eicosane content in such a way that a high n-eicosane content improved the microencapsulation efficiency, especially at higher SF/CS ratios.
Phadungphatthanakoon et al. [48] encapsulated n-eicosane (C20) into a blend of methyl cellulose (MC) and ethyl cellulose (EC). They observed up to a 24 and 29% increase in the heat of phase transition during crystallization and melting for the microparticles (C20/EC/MC with a 9% EC/MC blend content), respectively. Next, C20-loaded EC/MC microparticles were incorporated into natural rubber latex to obtain natural rubber-based composites with an acceptable thermoregulation property and improved mechanical properties.
In another development, PCM microcapsules based on n-nonadecane as the core material and sodium alginate (SA) as the shell material were fabricated by Moghaddam et al. [6] using a melt coaxial electrospray. It has been observed that the microcapsule size was increased with the SA concentration, while smaller microcapsule diameters and a more homogenous dimensional distribution were found for SA concentrations of 1.5% (w/v). A decrease in the needle-to-collector distance from 20 to 5 cm caused mean microcapsule size decrease from 480 to 275 µm. A further reduction in the working distance up to 3 cm allowed them to obtain spherically shaped microcapsules with diameters lower than 100 µm. The obtained results confirmed that for microcapsules with diameters smaller than 100 µm, the n-nonadecane encapsulation ratio was 56 ± 5%.
Fashandi and co-workers [54] investigated microencapsulation process of palmitic acid (PA) of vegetable origin in bio-based PLA shells using solvent evaporation and oil-in-water emulsification methods. They revealed that higher PA content lead to higher core content, and an increase in microPCM’s size. This was connected to the higher viscosity of the oil phase. The obtained results exhibited that PVA is a good emulsifier, and that proper PVA concentration choice make it possible to control the average size of microPCM capsules.
Multifunctional magnetic microcapsules based on n-eicosan and chitosan were fabricated by Chen et al. [24]. They used multi-emulsification and crosslinking approaches, as showed in Figure 4. The obtained microcapsules exhibit a regular spherical morphology with a uniform diameter of 500 μm and a heat of phase transition of up to 80 J/g.
Tangsiriratana et al. [56] studied microencapsulated PCMs composed of gelatin/gum Arabic (shell) and a sugarcane wax–Al2O3 composite (core), obtained via complex coacervation. The authors found that the heat storage-dissipation performance and thermal stability of PV panels with the sugarcane wax-based PCM composite in solar panels was influenced by its thickness. Additionally, the PV panels’ front-facing glass peak temperature was reduced by ca. 4% when the layer thickness of the encapsulated PCM was increased from 4 to 7 mm.
Reddy et al. [46] obtained PCMs based on caprylic acid, hexadecane, and butyl stearate encapsulated in the barium-crosslinked pectin shell following the ionic gelation protocol as a core—Figure 5.
Pectin/PCM capsules with diameters lower than 2 mm were obtained with an encapsulation efficiency of 83.66 wt.%, 83.21 wt.%, and 84.39 wt.% for hexadecane, butyl stearate, and caprylic acid as the PCM, respectively. They were characterized by melting enthalpies of 184.89 kJ/kg, 116 kJ/kg, and 118 kJ/kg. It was also revealed that pectin encapsulation caused an improvement in the thermal stability of encapsulated PCMs by 70–130 °C, compared to pristine PCMs.

5. PCMs Shape Stabilization Using Bio-Based Polymers

Another method to avoid the leakage of PCMs above their melting temperature is shape stabilization, usually by using porous materials. The shape-stabilization method with porous lightweight materials can limit PCM leakage during the phase transition, but can also improve selected properties such as the thermal conductivity [79].
PEO blends with selected polymers of natural origin or their derivatives—cellulose, cellulose acetate, cellulose ether, carboxymethoxycellulose, and potato starch—were obtained and characterized by the DSC and IR techniques. It has been shown that the phase transformations of the solid–solid type occur in PEO/natural polymer systems and, in the case of selected blends (PEO/cellulose acetate 1:1 w/w and PEO/CMC 1:1 w/w), a synergistic effect appears; however, the obtained systems are characterized by a relatively low phase-transition heat. The shape-stabilization effect was attributed to intermolecular hydrogen interactions between ether oxygen in PEO chains and cellulose (or its derivative) hydroxyl groups, or PEO hydroxyl end groups and oxygen atoms in cellulose macrochains [7,40].
Zhao et al. [50] prepared n-octadecane/silk composite fibers designed as form-stable PCMs by the emulsion–electrospinning method. Microscopic observations showed that the composite fibers are characterized by a cylindrical shape, uniform diameter, and smooth surface, while the DSC results revealed a high thermal energy storage capacity with reversible phase-transition behavior and a good thermal reliability after thermal cycling.
A eutectic mixture of TD and MA was incorporated into hydroxylpropyl methyl cellulose (HPMC) by Qu et al. [45]. The obtained results confirmed that a TD–MA eutectic mixture was distributed uniformly in HPMC, and a form-stable PCMs were obtained. Leakage tests revealed that the absorption efficiency of TD-MA by HPMC was nearly 50%. The thermal properties investigations showed that the temperatures of melting and freezing were 34.61 and 31.09 °C, and the heat of phase transition was 102.11 and 84.58 J/g, respectively. Importantly, the TD-MA/HPMC exhibited a good thermal stability and reliability performance.
Liu et al. [34] reported on shape-stabilized PCMs based on PEG 6000 and SA, prepared via a sol–gel method. Ca2+-crosslinked sodium alginate was modified with PEG—Figure 6. The PEG loading rate in the PEG/SA composites (PSCs) was 93% with a phase-transition enthalpy of up to 156.8 J/g and a satisfactory shape stability.
PEG-based PCMs shape-stabilized by chitin nanofibers (CNFs) were obtained by Wijesena et al. [23]. Fibrous chitin from crab shells was characterized by a diameter of several tens of nanometers and lengths of up to few micrometers. The best system remained opaque (transmittance ∼3.5%) below the melting temperature of PEG, while above the melting temperature it was transparent with a transmittance value of ∼88%.
Pinto and co-workers [64] prepared shape-stabilized PCMs based on bacterial cellulose nanofibers (BC) with reduced and non-reduced graphene oxide (GO) and microencapsulated paraffin as PCMs in the foam-like form. They revealed that the presence of GO enhanced the fire-retardancy and dimensional stability of the obtained system.
Capric acid and lauric acid encapsulated in SA were integrated with oakum fiber in a plaster matrix by M’ghari et al. [55]. The obtained results confirmed a good compatibility between the components and a sufficient chemical and thermal stability; a phase-change temperature of 16.54 °C and heat of phase transition of 35.18 J/g were determined after thermal cycling. The designed composite material can be considered, due to the high heat capacity, low cost, and proper phase-transition temperature, for energy storage in the building sector.
Bio-based poly(glycerol-itaconic acid) (PGI) as a support for PEG-based form-stable PCMs was proposed and tested by Yin et al. [35]. The PGI/PEG precursor has been synthesized by solvent-free polycondensation (Figure 7), where PEG as a PCM was encapsulated within the crosslinked PGI. Interestingly, PGI can entrap up to 72.67% of PEG 6000 with an excellent form stability and phase-transition enthalpies of up to 61.8–86.9 J/g. The form stability of the PGI/PEG system was supported by intermolecular hydrogen bonds, and the phase-change temperature and heat of PCMs can be controlled by both the PEG load and its average molar mass.
Ahn et al. [62] fabricated polymer fibers with a microencapsulated PCMs load up to 80 wt.% by incorporating commercially available PCM microcapsules into cellulose acetate, polyethersulfone, or cellulose and processing by solution-spinning. The obtained fibers were characterized by ca. 95% of thermal storage capacity, compared to the microcapsules used. Importantly, the DMA results show some relationships between the mechanical stability of PCM-polymer fibers and PCM load.
PEG2000–10,000 was impregnated into CNT/cotton yarn (CCY) and coated by electrospun PAN [36]. The obtained yarn contained conductive yarn with PEG located between the cotton fibers, and it was characterized by a heat of phase transition in the range of 126–150 J/g.
Singh et al. [42] proposed to use PCM beads to provide thermal buffering to a chocolate-containing box subjected to higher temperatures. For this purpose, 1-dodecanol was incorporated into barium alginate beads by ionotropic gelation and coated with a polyurea shell. The latent heat of the microcapsules without the polyurea coating was ca. 149 J/g, while after polyurea coating had been applied it amounted to ca. 76 J/g. The obtained beads were introduced to the inner walls of the packaging box to keep the temperature around 25 °C; the performed tests showed that the developed capsules could maintain the chocolate temperature at 25 °C by 86 min.
Meng et al. [59] used sodium alginate for the stabilization of sodium sulfate decahydrate (Na2SO4·10H2O) (SSD); the SA/SSD systems were prepared by blending SSD with SA at different concentrations. The best composite PCMs exhibited a phase-change enthalpy of up to 160 J/g, and the stabilization effect was attributed to the ionic and hydrogen bond interactions between SA and SSD.
Electrospun phase-change nanofiber films based on lignin were obtained by electrospinning by Niu et al. [26] using sodium lignosulfonate (SLS), poly(vinyl alcohol) (PVA), and PEG. In the proposed system, the SLS/PVA mixture acted as a supporting matrix; additionally, SLS is a photothermal material and interacts with PEG via hydrogen bonds and electrostatic forces. The obtained nanofibers exhibited a good solar-to-thermal energy conversion and storage ability with a latent enthalpy of fusion of 42.16 J/g, whereby the solar-to-thermal energy conversion and storage efficiency amounted to ca. 18%.
Wang et al. [52] studied a composite PCM based on silk fiber (SF) for wearable personal thermal management [52]. For preparing the PCM composite, natural silkworm cocoons and capric acid (CA) were used. The performed investigations revealed that SF and CA were physically bonded, and the crystal structure of CA was not changed by SF. The heat and temperature of the phase transition of SF/CA were 123.4 J/g and 30.6 °C, respectively, and the SF/CA composite showed a good shape and thermal stability.
Lignin–g–poly(ε-caprolactone) as a form-stable PCM was investigated by Lee et al. [60]. The copolymer was synthesized by grafting of PCL with alkaline lignin via ring-opening polymerization, while the physical mixtures of lignin with PCL were also prepared for comparative study. DSC data revealed that the melting temperature of lignin–g–PCL was ca. 51.3 °C and the latent enthalpy of phase transition was ca. 61 J/g. Tests also revealed its good thermal reliability and stability after a thermal cycling test of up to 100 cycles.
Liu and co-workers [41] prepared and tested coaxial fiber membranes (CFMs) with PEG stabilized by bio-based PLLA, prepared using coaxial electrospinning. The performed studies indicated that CFMs with a PEG content of 49 wt.% show a good energy storage ability and thermal stability. Moreover, in comparison with the PLLA membrane without PEG, the thermal insulation efficiency was improved more than 21%.
In another work, Li et al. [25] obtained porous capsules as shape-stable PCMs based on PEG, and a cellulose polyelectrolyte complex synthesized using carboxymethyl cellulose (anionic polyelectrolyte) and cellulose 6-(N-pyridinium)hexanoyl ester (cationic polyelectrolyte). To improve the thermal conductivity and photothermal conversion efficiency, MWCNTs or GO were introduced. Capsules with a core-shell structure exhibited a thermal energy storage ability with the latent enthalpy of phase transition 142.2 J/g and promising form-stable properties, thermal stability, and photothermal conversion performance. Erythritol (ET)-grafted polylactic acid (PLA) (ETPLA) polymers were in situ synthesized as shape-stabilized PCMs by Lee et al. [61]. The thermal conductivity of ETPLA was improved by modification with an oxidized carbon fiber (CF-OH) and aluminum nitride (AlN). The obtained PCM composite exhibited a thermal conductivity of 4.25 W/mK, much better mechanical properties compared to pure ET, and a very high latent enthalpy of fusion of up to 170.7 J/g, making the fabricated PCM composites a good candidate for thermal management systems of electronic devices.
Baniasadi et al. [27] used PCL for the shape stabilization of PEG nanofiber mats obtained by electrospinning. Additionally, gelatin and curcumin were applied to enhance the biomedical performance of the fabricated textiles. Microstructure analysis of the mats indicated nanofibers with randomly oriented morphology and diameters in the range of 220–270 nm. Mats were also characterized by good water absorption and mechanical properties. The heat of phase transition was 61.7 J/g with a good thermal reliability over 100 heating–cooling cycles. Moreover, the textiles modified with curcumin show a significant antioxidant activity, indicating a great potential application as a biomedical dressing that reduces inflammation and exerts a cooling effect.
Highly porous PEG-based materials with modified cellulose nanofibrils (CNF), lignin-containing CNF (LCNF), or acetylated CNF (ACNF), containing titanium dioxide, were prepared by Le et al. [32]—Figure 8. The PEG/CNF hybrid was characterized by a very low density in the range of 0.022–0.043 g/cm3 and a high latent enthalpy of phase transition of up to 204 J/g. The obtained system was tested as a material for buildings’ thermal management, and it was found that under excess solar heating, the PCM composite provided a good insulative protection and exceptional acoustic absorbance; modification with titanium dioxide nanoparticles improved the thermoregulating abilities.
In PEG20,000/PLA/graphitic carbon nitride (g-C3N4) composite PCM, prepared using the confinement method, PEG hinders the crystallization of PLA, and it was confirmed using the XRD method. The authors found that the incorporation of PEG changed the number of diffraction peaks of PLA, suggesting disrupted PLA crystallization, and in consequence a decrease in hardness [29]. The composite PCM exhibited a heat of phase transition of 106.1 kJ/kg for PLA:PEG:g-C3N4 with a mass ratio 3:6:1, and an increased thermal conductivity compared to PLA.
Kizildag [65] used microcapsules with melamine–formaldehyde shells and paraffin cores as PCMs incorporated into pullulan solution. Pullulan films were obtained by the film-casting method of prepared dispersion (75 wt.% of PCM), and they showed an enthalpy of phase transition of 104.85 J/g during heating and 103.58 J/g during cooling, as well as a good thermal reliability after thermal cycling.

6. Conclusions

Bio-based polymers are currently considered as future candidates for the design and manufacturing of modern phase-change materials. The application of bio-based polymeric materials allows us to minimize PCMs’ environmental impact and carbon footprint, especially in the energy- and materials-consuming building and energy sectors. Even a partial replacement of synthetic polymers with bio-based ones makes it possible to decrease the negative impact of PCMs on the natural environment and carbon footprint. Current research works focus on two topics: (i) PCMs encapsulation using bio-based polymers, and (ii) PCMs shape stabilization by bio-based polymers. In the former area, to avoid leakage, as well as to protect and enhance PCM properties, various encapsulation strategies have been developed, and the most up-to-date trend is the use of bio-based components in the (micro)capsules preparation process. It has been found that the application of bio-based polymers, for instance cellulose and its derivatives, enables the fabrication of encapsulated PCMs with good thermoregulation abilities and acceptable mechanical properties. Interestingly, microencapsulated PCMs composed of gelatin/gum Arabic (polymer shell), and sugarcane wax–Al2O3 composite (core material) have already been tested in solar panels, and a desired reduction in the front-facing glass temperature was observed. In the second area, i.e., PCMs shape stabilization, bio-based polymers are applied to limit PCM leakage during the phase transition, but can also improve selected properties, such as thermal conductivity, as well as shape stability which is of primary importance. In synthetic/natural polymer blends, e.g., PEO/cellulose (derivatives), intermolecular hydrogen interactions between oxygen atoms in PEO chains and hydroxyl groups in cellulose chains (or its derivatives), or hydroxyl end groups in PEO chains and oxygen atoms in cellulose chains, were found to improve the shape stabilization effect. On the other hand, the physical modification of PCMs has also been performed to impart new features to the shape-stabilized systems; e.g., the incorporation of graphene oxide remarkable improved the fire-retardancy of bacterial cellulose-based PCM. In another development, PCMs modified by multi-walled carbon nanotubes were found to provide an advantageous light-to-thermal conversion performance.
To conclude, bio-based polymers are a vast group of materials that can be applied in the design and manufacturing of environmentally friendly phase-change materials with a reduced carbon footprint. Biopolymers can be chemically or physically modified to give them new features, and such new (composite) materials will gradually replace components produced from non-renewable sources and thus contribute to the development of a future advanced ‘green’ economy.

Funding

The authors (K.P. (Kinga Pielichowska), K.N.D.) are grateful to the Polish National Science Centre for financial support under Contract No. UMO-2016/22/E/ST8/00048. This work was supported from the subsidy of the Ministry of Education and Science for the AGH University of Science and Technology in Kraków (Project No 16.16.160.557). The research project was supported/partly supported by program “Excellence initiative—research university” for the AGH University of Krakow.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

BCbacterial cellulose nanofibers
CAcapric acid
CACcellulose acetate
CCYcotton yarn
CELcellulose
CETmethylhydroxycellulose
CMCcarboxymethylcellulose
CNFschitin nanofibers
CNTcarbon nanotubes
CSchitosan
DCMdichloromethane
DSCdifferential scanning calorimetry
ECethyl cellulose
GOgraphene oxide
HPMChydroxylpropyl methyl cellulose
LAlauric acid
MAmyristic acid
MCmethyl cellulose
PApalmitic acid
PANpolyacrylonitrile
PBATpoly(butylene adipate-co-butylene terephthalate)
PBSApoly(butylene succinate-co-butylene adipate)
PCLpoly(ε-caprolactone)
PCMphase-change material
PDApolydopamine
PEFpoly(ethylene furanoate)
PEGpoly(ethylene glycol)
PEOpoly(ethylene oxide)
PETpoly(ethylene terephthalate)
PGIpoly(glycerol-itaconic acid)
PLApoly(lactic acid)
POMpolyoxymethylene
PPOpoly(propylene oxide)
PTHFpolytetrahydrofurane
PTTpoly(trimethylene terephthalate)
PVApoly(vinyl alcohol)
SFsilk fibroin
TDtetradecanol

References

  1. Pielichowska, K.; Pielichowski, K. Phase Change Materials for Thermal Energy Storage. Prog. Mater. Sci. 2014, 65, 67–123. [Google Scholar] [CrossRef]
  2. Zalba, B.; Marin, J.M.; Cabeza, L.F.; Mehling, H. Review on Thermal Energy Storage with Phase Change: Materials, Heat Transfer Analysis and Applications. Appl. Therm. Eng. 2003, 23, 251–283. [Google Scholar] [CrossRef]
  3. Pielichowska, K.; Pielichowski, K. (Eds.) Multifunctional Phase Change Materials; Elsevier: Amsterdam, The Netherlands, 2023; ISBN 9780323857192. [Google Scholar]
  4. Confalonieri, C.; Gariboldi, E. 9—Shape-Stabilized and Form-Stable PCMs. In Multifunctional Phase Change Materials; Pielichowska, K., Pielichowski, K., Eds.; Woodhead Publishing: Southton, UK, 2023; pp. 369–410. ISBN 978-0-323-85719-2. [Google Scholar]
  5. Pielichowska, K.; Bieda, J.; Szatkowski, P. Polyurethane/Graphite Nano-Platelet Composites for Thermal Energy Storage. Renew. Energy 2016, 91, 456–465. [Google Scholar] [CrossRef]
  6. Moghaddam, M.K.; Mortazavi, S.M.; Khayamian, T. Preparation of Calcium Alginate Microcapsules Containing N-Nonadecane by a Melt Coaxial Electrospray Method. J. Electrost. 2015, 73, 56–64. [Google Scholar] [CrossRef]
  7. Pielichowska, K.; Pielichowski, K. Novel Biodegradable Form Stable Phase Change Materials: Blends of Poly(Ethylene Oxide) and Gelatinized Potato Starch. J. Appl. Polym. Sci. 2010, 116, 1725–1731. [Google Scholar] [CrossRef]
  8. Okogeri, O.; Stathopoulos, V.N. What about Greener Phase Change Materials? A Review on Biobased Phase Change Materials for Thermal Energy Storage Applications. Int. J. Thermofluids 2021, 10, 100081. [Google Scholar] [CrossRef]
  9. Masutani, K.; Kimura, Y. Biobased Polymers BT—Encyclopedia of Polymeric Nanomaterials; Kobayashi, S., Müllen, K., Eds.; Springer: Berlin/Heidelberg, Germany, 2021; pp. 1–7. ISBN 978-3-642-36199-9. [Google Scholar]
  10. Piorkowska, E. Overview of Biobased Polymers. In Thermal Properties of Bio-Based Polymers; Di Lorenzo, M.L., Androsch, R., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 1–35. ISBN 978-3-030-39962-7. [Google Scholar]
  11. Nakajima, H.; Dijkstra, P.; Loos, K. The Recent Developments in Biobased Polymers toward General and Engineering Applications: Polymers That Are Upgraded from Biodegradable Polymers, Analogous to Petroleum-Derived Polymers, and Newly Developed. Polymers 2017, 9, 523. [Google Scholar] [CrossRef]
  12. Narayan, R. Preparation of Bio-Based Polymers for Materials Applications. Appl. Biochem. Biotechnol. 1988, 17, 7–22. [Google Scholar] [CrossRef]
  13. Carafa, R.N.; Foucher, D.A.; Sacripante, G.G. Biobased Polymers from Lignocellulosic Sources. Green Chem. Lett. Rev. 2023, 16, 2153087. [Google Scholar] [CrossRef]
  14. Kimura, Y. Molecular, Structural, and Material Design of Bio-Based Polymers. Polym. J. 2009, 41, 797–807. [Google Scholar] [CrossRef]
  15. Hatti-Kaul, R.; Nilsson, L.J.; Zhang, B.; Rehnberg, N.; Lundmark, S. Designing Biobased Recyclable Polymers for Plastics. Trends Biotechnol. 2020, 38, 50–67. [Google Scholar] [CrossRef]
  16. Mtibe, A.; Motloung, M.P.; Bandyopadhyay, J.; Ray, S.S. Synthetic Biopolymers and Their Composites: Advantages and Limitations—An Overview. Macromol. Rapid Commun. 2021, 42, 2100130. [Google Scholar] [CrossRef] [PubMed]
  17. Available online: https://plasticseurope.org/knowledge-hub/plastics-the-facts-2022/ (accessed on 20 November 2023).
  18. Available online: https://www.cas.org/resources/cas-insights/materials/biopolymers-manufacturings-latest-green-hero (accessed on 20 November 2023).
  19. Available online: https://www.european-bioplastics.org/market/ (accessed on 20 November 2023).
  20. Available online: https://bioplasticsnews.com/top-bioplastics-producers/ (accessed on 20 November 2023).
  21. Available online: https://biokunststofftool.de/materials/bio-pp/?lang=en (accessed on 20 November 2023).
  22. Available online: https://natureplast.eu/en/matiere/biobased-pa-2/ (accessed on 20 November 2023).
  23. Wijesena, R.N.; Tissera, N.D.; Rathnayaka, V.W.S.G.; Rajapakse, H.D.; de Silva, R.M.; de Silva, K.M.N. Shape-Stabilization of Polyethylene Glycol Phase Change Materials with Chitin Nanofibers for Applications in “Smart” Windows. Carbohydr. Polym. 2020, 237, 116132. [Google Scholar] [CrossRef] [PubMed]
  24. Chen, X.; Zhao, Y.; Zhang, Y.; Lu, A.; Li, X.; Liu, L.; Qin, G.; Fang, Z.; Zhang, J.; Liu, Y. A Novel Design and Synthesis of Multifunctional Magnetic Chitosan Microsphere Based on Phase Change Materials. Mater. Lett. 2019, 237, 185–187. [Google Scholar] [CrossRef]
  25. Li, J.; Meng, L.; Chen, J.; Chen, X.; Wang, Y.; Xiao, Z.; Wang, H.; Liang, D.; Xie, Y. Encapsulation of Polyethylene Glycol in Cellulose-Based Porous Capsules for Latent Heat Storage and Light-to-Thermal Conversion. Front. Chem. Sci. Eng. 2023, 17, 1038–1050. [Google Scholar] [CrossRef]
  26. Niu, L.; Li, X.; Zhang, Y.; Yang, H.; Feng, J.; Liu, Z. Electrospun Lignin-Based Phase-Change Nanofiber Films for Solar Energy Storage. ACS Sustain. Chem. Eng. 2022, 10, 13081–13090. [Google Scholar] [CrossRef]
  27. Baniasadi, H.; Madani, M.; Seppälä, J.; Zimmerman, J.B.; Yazdani, M.R. Form-Stable Phase Change Electrospun Nanofibers Mat with Thermal Regulation and Biomedical Multi-Functionalities. J. Energy Storage 2023, 68, 107660. [Google Scholar] [CrossRef]
  28. Li, Y.-Q.; Huang, X.; Li, Y.; Xi, Z.; Hai, G.; Tao, Z.; Wang, G. Shape-Stabilized Phase-Change Materials Supported by Eggplant-Derived Porous Carbon for Efficient Solar-to-Thermal Energy Conversion and Storage. Sustain. Energy Fuels 2020, 4, 1764–1772. [Google Scholar] [CrossRef]
  29. Feng, L.; Ding, J.; Hu, H.; Lv, Z.; Zhang, Y.; Xu, B.; Quan, J.; Hao, S.; Fan, H.; Hang, Z. Preparation and Characterization of Bio-Based PLA/PEG/g-C3N4 Low-Temperature Composite Phase Change Energy Storage Materials. Polymers 2023, 15, 2872. [Google Scholar] [CrossRef]
  30. Sheng, X.; Dong, D.; Lu, X.; Zhang, L.; Chen, Y. MXene-Wrapped Bio-Based Pomelo Peel Foam/Polyethylene Glycol Composite Phase Change Material with Enhanced Light-to-Thermal Conversion Efficiency, Thermal Energy Storage Capability and Thermal Conductivity. Compos. Part A Appl. Sci. Manuf. 2020, 138, 106067. [Google Scholar] [CrossRef]
  31. Wei, D.; Wu, C.; Jiang, G.; Sheng, X.; Xie, Y. Lignin-Assisted Construction of Well-Defined 3D Graphene Aerogel/PEG Form-Stable Phase Change Composites towards Efficient Solar Thermal Energy Storage. Sol. Energy Mater. Sol. Cells 2021, 224, 111013. [Google Scholar] [CrossRef]
  32. Le, W.T.; Kankkunen, A.; Rojas, O.J.; Yazdani, M.R. Leakage-Free Porous Cellulose-Based Phase Change Cryogels for Sound and Thermal Insulation. Sol. Energy Mater. Sol. Cells 2023, 256, 112337. [Google Scholar] [CrossRef]
  33. Baniasadi, H.; Seppälä, J.; Kankkunen, A.; Seppälä, A.; Yazdani, M.R. Water-Resistant Gum-Based Phase Change Composite for Thermo-Regulating Insulation Packaging. J. Energy Storage 2023, 61, 106725. [Google Scholar] [CrossRef]
  34. Liu, L.; Fan, X.; Zhang, Y.; Zhang, S.; Wang, W.; Jin, X.; Tang, B. Novel Bio-Based Phase Change Materials with High Enthalpy for Thermal Energy Storage. Appl. Energy 2020, 268, 114979. [Google Scholar] [CrossRef]
  35. Yin, G.Z.; Díaz Palencia, J.L.; Wang, D.Y. Fully Bio-Based Poly (Glycerol-Itaconic Acid) as Supporter for PEG Based Form Stable Phase Change Materials. Compos. Commun. 2021, 27, 100893. [Google Scholar] [CrossRef]
  36. Ke, G.; Jin, X.; Cai, G.; Li, W.; Xu, A. A Novel Composite Cotton Yarn with Phase Change and Electrical Conductivity Functions. J. Ind. Text. 2022, 51, 554S–568S. [Google Scholar] [CrossRef]
  37. Xie, Y.; Li, W.; Huang, H.; Dong, D.; Zhang, X.; Zhang, L.; Chen, Y.; Sheng, X.; Lu, X. Bio-Based Radish@PDA/PEG Sandwich Composite with High Efficiency Solar Thermal Energy Storage. ACS Sustain. Chem. Eng. 2020, 8, 8448–8457. [Google Scholar] [CrossRef]
  38. Chen, C.; Zhao, Y.; Liu, W. Electrospun Polyethylene Glycol/Cellulose Acetate Phase Change Fibers with Core-Sheath Structure for Thermal Energy Storage. Renew. Energy 2013, 60, 222–225. [Google Scholar] [CrossRef]
  39. Zhang, H.-C.; Kang, B.-h.; Sheng, X.; Lu, X. Novel Bio-Based Pomelo Peel Flour/Polyethylene Glycol Composite Phase Change Material for Thermal Energy Storage. Polymers 2019, 11, 2043. [Google Scholar] [CrossRef]
  40. Pielichowska, K.; Pielichowski, K. Biodegradable PEO/Cellulose-Based Solid-Solid Phase Change Materials. Polym. Adv. Technol. 2011, 22, 1633–1641. [Google Scholar] [CrossRef]
  41. Liu, Y.; Wu, Y.; Zhao, S.; Wang, X.; Zheng, J.; Zeng, W.; Yuan, M.; Zhao, N.; Li, Q.; Wang, Z.; et al. Biobased Phase Change Material with Reduced Thermal Conductivity: From Preparation to Analysis of Thermal Insulation Performance. ACS Appl. Polym. Mater. 2023, 5, 3728–3736. [Google Scholar] [CrossRef]
  42. Singh, J.; Vennapusa, J.R.; Dixit, P.; Maiti, T.K.; Chattopadhyay, S. A Novel Strategy for Temperature Controlling of Chocolates through 1-Dodecanol Embedded Polyurea Coated Barium Alginate Beads. J. Taiwan Inst. Chem. Eng. 2022, 138, 104497. [Google Scholar] [CrossRef]
  43. Atinafu, D.G.; Yang, S.; Yun, B.Y.; Kang, Y.; Kim, S. Use of Biochar Co-Mediated Chitosan Mesopores to Encapsulate Alkane and Improve Thermal Properties. Environ. Res. 2022, 212, 113539. [Google Scholar] [CrossRef] [PubMed]
  44. Montanari, C.; Chen, H.; Lidfeldt, M.; Gunnarsson, J.; Olsén, P.; Berglund, L.A. Sustainable Thermal Energy Batteries from Fully Bio-Based Transparent Wood. Small 2023, 19, 2301262. [Google Scholar] [CrossRef]
  45. Qu, M.; Guo, C.; Li, L.; Zhang, X. Preparation and Investigation on Tetradecanol and Myristic Acid/Cellulose Form-Stable Phase Change Material. J. Therm. Anal. Calorim. 2017, 130, 781–790. [Google Scholar] [CrossRef]
  46. Reddy, V.J.; Dixit, P.; Singh, J.; Chattopadhyay, S. Understanding the Core-Shell Interactions in Macrocapsules of Organic Phase Change Materials and Polysaccharide Shell. Carbohydr. Polym. 2022, 294, 119786. [Google Scholar] [CrossRef]
  47. Atinafu, D.G.; Yeol Yun, B.; Uk Kim, Y.; Wi, S.; Kim, S. Introduction of Eicosane into Biochar Derived from Softwood and Wheat Straw: Influence of Porous Structure and Surface Chemistry. Chem. Eng. J. 2021, 415, 128887. [Google Scholar] [CrossRef]
  48. Phadungphatthanakoon, S.; Poompradub, S.; Wanichwecharungruang, S.P. Increasing the Thermal Storage Capacity of a Phase Change Material by Encapsulation: Preparation and Application in Natural Rubber. ACS Appl. Mater. Interfaces 2011, 3, 3691–3696. [Google Scholar] [CrossRef]
  49. Jin, L.; Tan, Y.; Yuan, S.; Wang, S.; Cheng, X.; Wang, H.; Du, Z.; Du, X. Phytic Acid–Decorated κ-Carrageenan/Melanin Hybrid Aerogels Supported Phase Change Composites with Excellent Photothermal Conversion Efficiency and Flame Retardancy. Renew. Energy 2023, 206, 148–156. [Google Scholar] [CrossRef]
  50. Zhao, L.; Luo, J.; Li, Y.; Wang, H.; Song, G.; Tang, G. Emulsion-Electrospinning n-Octadecane/Silk Composite Fiber as Environmental-Friendly Form-Stable Phase Change Materials. J. Appl. Polym. Sci. 2017, 134, 45538. [Google Scholar] [CrossRef]
  51. Baştürk, E.; Kahraman, M.V. Photocrosslinked Biobased Phase Change Material for Thermal Energy Storage. J. Appl. Polym. Sci. 2016, 133, 43757. [Google Scholar] [CrossRef]
  52. Wang, C.; Dong, H.; Cheng, C.; Sun, K.; Jin, T.; Shi, Q. Flexible and Biocompatible Silk Fiber-Based Composite Phase Change Material for Personal Thermal Management. ACS Sustain. Chem. Eng. 2022, 10, 16368–16376. [Google Scholar] [CrossRef]
  53. Mandal, S.; Ishak, S.; Lee, D.-E.; Park, T. Shape-Stabilized Orange Peel/Myristic Acid Phase Change Materials for Efficient Thermal Energy Storage Application. Energy Rep. 2022, 8, 9618–9628. [Google Scholar] [CrossRef]
  54. Fashandi, M.; Leung, S.N. Preparation and Characterization of 100% Bio-Based Polylactic Acid/Palmitic Acid Microcapsules for Thermal Energy Storage. Mater. Renew. Sustain. Energy 2017, 6, 14. [Google Scholar] [CrossRef]
  55. M’ghari, O.; Hassani, F.S.A.; Mekhzoum, M.E.M.; Zari, N.; Bouhfid, R.; el Kacem Qaiss, A. Elaboration of a Composite Material Based on Plaster Reinforced with Phase Change Material/Oakum Fiber: Physical, Thermal and Mechanical Properties. J. Energy Storage 2021, 35, 102321. [Google Scholar] [CrossRef]
  56. Tangsiriratana, E.; Skolpap, W.; Patterson, R.J.; Sriprapha, K. Thermal Properties and Behavior of Microencapsulated Sugarcane Wax Phase Change Material. Heliyon 2019, 5, e02184. [Google Scholar] [CrossRef] [PubMed]
  57. Abdulmunem, A.R.; Samin, P.M.; Sopian, K.; Hoseinzadeh, S.; Al-Jaber, H.A.; Garcia, D.A. Waste Chicken Feathers Integrated with Phase Change Materials as New Inner Insulation Envelope for Buildings. J. Energy Storage 2022, 56, 106130. [Google Scholar] [CrossRef]
  58. Souissi, M.; Trigui, A.; Jedidi, I.; Loukil, M.S.; Abdelmouleh, M. Bio-Based Composite as Phase Change Material Including Spent Coffee Grounds and Beeswax Paraffin. Korean J. Chem. Eng. 2023, 40, 2342–2355. [Google Scholar] [CrossRef]
  59. Meng, L.; Ivanov, A.S.; Kim, S.; Zhao, X.; Kumar, N.; Young-Gonzales, A.; Saito, T.; Bras, W.; Gluesenkamp, K.; Bocharova, V. Alginate-Sodium Sulfate Decahydrate Phase Change Composite with Extended Stability. ACS Appl. Polym. Mater. 2022, 4, 6563–6571. [Google Scholar] [CrossRef]
  60. Lee, J.J.C.; Sugiarto, S.; Ong, P.J.; Soo, X.Y.D.; Ni, X.; Luo, P.; Hnin, Y.Y.K.; See, J.S.Y.; Wei, F.; Zheng, R.; et al. Lignin-g-Polycaprolactone as a Form-Stable Phase Change Material for Thermal Energy Storage Application. J. Energy Storage 2022, 56, 106118. [Google Scholar] [CrossRef]
  61. Lee, W.; Lee, J.; Yang, W.; Kim, J. Fabrication of Biobased Advanced Phase Change Material and Multifunctional Composites for Efficient Thermal Management. ACS Sustain. Chem. Eng. 2023, 11, 1178–1189. [Google Scholar] [CrossRef]
  62. Ahn, Y.-H.; DeWitt, S.J.A.; McGuire, S.; Lively, R.P. Incorporation of Phase Change Materials into Fibers for Sustainable Thermal Energy Storage. Ind. Eng. Chem. Res. 2021, 60, 3374–3384. [Google Scholar] [CrossRef]
  63. Oliveira, F.R.; Fernandes, M.; Carneiro, N.; Pedro Souto, A. Functionalization of Wool Fabric with Phase-Change Materials Microcapsules after Plasma Surface Modification. J. Appl. Polym. Sci. 2013, 128, 2638–2647. [Google Scholar] [CrossRef]
  64. Pinto, S.C.; Silva, N.H.C.S.; Pinto, R.J.B.; Freire, C.S.R.; Duarte, I.; Vicente, R.; Vesenjak, M.; Marques, P.A.A.P. Multifunctional Hybrid Structures Made of Open-Cell Aluminum Foam Impregnated with Cellulose/Graphene Nanocomposites. Carbohydr. Polym. 2020, 238, 116197. [Google Scholar] [CrossRef] [PubMed]
  65. Kizildag, N. Pullulan Films with PCMs: Recyclable Bio-Based Films with Thermal Management Functionality. Coatings 2023, 13, 414. [Google Scholar] [CrossRef]
  66. Kamimoto, M.; Abe, Y.; Sawata, S.; Tani, T.; Ozawa, T. Latent Thermal Storage Unit Using Form-Stable High Density Polyethylene; Part I: Performance of the Storage Unit. J. Sol. Energy Eng. 1986, 108, 282–289. [Google Scholar] [CrossRef]
  67. Weingrill, H.M.; Resch-Fauster, K.; Lucyshyn, T.; Zauner, C. High-Density Polyethylene as Phase-Change Material: Long-Term Stability and Aging. Polym. Test. 2019, 76, 433–442. [Google Scholar] [CrossRef]
  68. Weingrill, H.M.; Resch-Fauster, K.; Zauner, C. Applicability of Polymeric Materials as Phase Change Materials. Macromol. Mater. Eng. 2018, 303, 1800355. [Google Scholar] [CrossRef]
  69. Iwamoto, Y.; Ikai, S. New Polymeric Material for Latent Heat Thermal Energy Storage. In Proceedings of the 5th Workshop of the IEA ECES IA Annex 10, Tsu, Japan, 10 November 2000. [Google Scholar]
  70. Herzberger, J.; Niederer, K.; Pohlit, H.; Seiwert, J.; Worm, M.; Wurm, F.R.; Frey, H. Polymerization of Ethylene Oxide, Propylene Oxide, and Other Alkylene Oxides: Synthesis, Novel Polymer Architectures, and Bioconjugation. Chem. Rev. 2016, 116, 2170–2243. [Google Scholar] [CrossRef]
  71. Bailey, F.E.; Koleske, J.V. Poly(Ethylene Oxide); Academic Press: Cambridge, MA, USA, 1976; ISBN 9780120732500. [Google Scholar]
  72. Pielichowski, K.; Flejtuch, K. Differential Scanning Calorimetry Study of Blends of Poly(Ethylene Glycol) with Selected Fatty Acids. Macromol. Mater. Eng. 2003, 288, 259–264. [Google Scholar] [CrossRef]
  73. Pielichowski, K.; Flejtuch, K. Differential Scanning Calorimetry Studies on Poly(Ethylene Glycol) with Different Molecular Weights for Thermal Energy Storage Materials. Polym. Adv. Technol. 2002, 13, 690–696. [Google Scholar] [CrossRef]
  74. Flejtuch, K. Badanie Przemian Fazowych Wybranych Ukladów Polieterów Pod Kątem Akumulacji Energii Cieplnej. PhD Thesis, Cracow University of Technology, Krakow, Poland, 2004. [Google Scholar]
  75. Pielichowski, K.; Flejtuch, K. Binary Blends of Polyethers with Fatty Acids: A Thermal Characterization of the Phase Transitions. J. Appl. Polym. Sci. 2003, 90, 861–870. [Google Scholar] [CrossRef]
  76. Huang, Y.; Stonehouse, A.; Abeykoon, C. Encapsulation Methods for Phase Change Materials—A Critical Review. Int. J. Heat Mass Transf. 2023, 200, 123458. [Google Scholar] [CrossRef]
  77. Ghasemi, K.; Tasnim, S.; Mahmud, S. PCM, Nano/Microencapsulation and Slurries: A Review of Fundamentals, Categories, Fabrication, Numerical Models and Applications. Sustain. Energy Technol. Assess. 2022, 52, 102084. [Google Scholar] [CrossRef]
  78. Deveci, S.S.; Basal, G. Preparation of PCM Microcapsules by Complex Coacervation of Silk Fibroin and Chitosan. Colloid Polym. Sci. 2009, 287, 1455–1467. [Google Scholar] [CrossRef]
  79. Jeon, I.K.; Azzam, A.; Al Jebaei, H.; Kim, Y.-R.; Aryal, A.; Baltazar, J.-C. Effects of Shape-Stabilized Phase Change Materials in Cementitious Composites on Thermal-Mechanical Properties and Economic Benefits. Appl. Therm. Eng. 2023, 219, 119444. [Google Scholar] [CrossRef]
Figure 1. Presentation of the production of bio-based plastics and their recycling. Reprinted from [15] with permission from Elsevier.
Figure 1. Presentation of the production of bio-based plastics and their recycling. Reprinted from [15] with permission from Elsevier.
Polymers 16 00328 g001
Figure 2. Classification of biopolymers. Reprinted from [16] with permission from Wiley.
Figure 2. Classification of biopolymers. Reprinted from [16] with permission from Wiley.
Polymers 16 00328 g002
Figure 3. Overview of the melting behavior of examined semi-crystalline polymers generated via DSC when priorly cooled at −10 °C min−1 and heated at 10 °C min−1. The melting temperature range of each polymer type reflects the temperature range between the lowest TMon and the higher TMend of any polymer of the according polymer type. Reprinted from [68] with permission from Wiley.
Figure 3. Overview of the melting behavior of examined semi-crystalline polymers generated via DSC when priorly cooled at −10 °C min−1 and heated at 10 °C min−1. The melting temperature range of each polymer type reflects the temperature range between the lowest TMon and the higher TMend of any polymer of the according polymer type. Reprinted from [68] with permission from Wiley.
Polymers 16 00328 g003
Figure 4. (a) SEM micrographs of Fe3O4 nanoparticles, (b) SEM micrographs of M-PCM-20-CTS, (c) high-magnification SEM micrographs of M-PCM-20-CTS, (d) the digital photos of an aqueous suspension containing M-PCM-20-CTS before and after magnetic separation. Reprinted from [24] with permission from Elsevier.
Figure 4. (a) SEM micrographs of Fe3O4 nanoparticles, (b) SEM micrographs of M-PCM-20-CTS, (c) high-magnification SEM micrographs of M-PCM-20-CTS, (d) the digital photos of an aqueous suspension containing M-PCM-20-CTS before and after magnetic separation. Reprinted from [24] with permission from Elsevier.
Polymers 16 00328 g004
Figure 5. Schematic diagram of a pectin–PCM capsules preparation method through ionic gelation. Reprinted from [46] with permission from Elsevier.
Figure 5. Schematic diagram of a pectin–PCM capsules preparation method through ionic gelation. Reprinted from [46] with permission from Elsevier.
Polymers 16 00328 g005
Figure 6. Schematic illustration of the PSCs synthesis process. Reprinted from [34] with permission from Elsevier.
Figure 6. Schematic illustration of the PSCs synthesis process. Reprinted from [34] with permission from Elsevier.
Polymers 16 00328 g006
Figure 7. (a) Synthesis of a PGI/PEG precursor by polycondensation (Step 1), (b) 13C NMR full spectrum of PGI/PEG (PGI-45) (before curing), (c) Schematic diagram of PGI/PEG precursor with free vinyl and -OH groups, (d) Cured PGI/PEG induced by the BPO initiator at 80 °C (Step 2) and (e) structure illustration of the reversible phase change. Reprinted from [35] with permission from Elsevier.
Figure 7. (a) Synthesis of a PGI/PEG precursor by polycondensation (Step 1), (b) 13C NMR full spectrum of PGI/PEG (PGI-45) (before curing), (c) Schematic diagram of PGI/PEG precursor with free vinyl and -OH groups, (d) Cured PGI/PEG induced by the BPO initiator at 80 °C (Step 2) and (e) structure illustration of the reversible phase change. Reprinted from [35] with permission from Elsevier.
Polymers 16 00328 g007
Figure 8. Form stability and leakage-proof phase transition of a phase-change nanohybrid (PCN-L3) composed of PEG (74 wt.%), lignin-CNF (24 wt.%), and TiO2 (2 wt.%) through intermolecular interactions and hydrogen bonding. Reprinted from [32] with permission from Elsevier.
Figure 8. Form stability and leakage-proof phase transition of a phase-change nanohybrid (PCN-L3) composed of PEG (74 wt.%), lignin-CNF (24 wt.%), and TiO2 (2 wt.%) through intermolecular interactions and hydrogen bonding. Reprinted from [32] with permission from Elsevier.
Polymers 16 00328 g008
Table 1. Bio-based polymers.
Table 1. Bio-based polymers.
BiopolymerBiobased Content (%)Major ProducersApplicationsBiodegradabilityRef.
Starch100Novamont; Futerro; BiomeFlexible packaging, consumer goods, agricultureYes[18,19]
PLAUp to 100NatureWorks; Total Corbion; Shimadzu Cor.; Toyobo; Evonik, Flexible packaging, rigid packaging, consumer goodsYes[18,19]
PHA100Bio- On; Kaneka; Tepha; Danimer Scientific; Newlight Technologies; Yield10 Bioscience, Tianjin GreenBio Materials, Flexible packaging, rigid packagingYes[18,19]
PHBn/aMitsubishi Chemical; Showa Denko K.K.; SK Chemicals; PTT MCC BiochemFlexible packaging, rigid packagingYes[19]
PBSTill 100Showa Denko K.K.; SK Chemicals; Mitsubishi ChemicalFlexible packaging, agricultureYes[18,19]
PBSTil 50Natureplast;Flexible packaging, rigid packaging, regular consumption goods, horticulture/agricultureYes[20]
PEGTil 100Sirius; Croda; Solvay; Clariant IGL Specialty Chemicals; BASF; Helvetic; BioscienceConsumer goodsYes[18]
PBSATill 100Showa Denko K.K.; SK Chemicals; Flexible packaging, agricultureYes[18,19]
PBATTill 50BASF; Zhuhai Wango Chemical Co.; JinHui
ZhaoLong; Eastman; Novamont
Flexible packaging, rigid packaging, agricultureYes[18,19]
PCLn/aPerstorp until 2018, acquired by IngevityConsumer goodsYes[18,19]
PTT27–37DupontFibers (including woven and non-woven), consumer goodsNo[18,19]
PEFTill 100Avantium; Corbion; BASF; DuPont, SynvinaPackaging: replacement for PETNo[18,19]
PEn/aBraskem; Neste; LyondellBasellFlexible packaging, rigid packaging, building and constructionNo[19,21]
PP30Borealis; Neste; LyondellBasellFlexible packaging, rigid packaging,No[21]
PATill 100Natureplast; Technoform; Nexis Fibers; Acro; Avient; Fulgar; Akeme; HengshuiTechnical parts, regular consumption goods, sports and leisure, transportNo [20,22]
PETTill 75Toray Industries, The Coca-Cola Company, M&G Chemicals; Virent; PepsiCo; Toyota TsushoRigid packagingNo[18,19]
Table 2. PCM systems with bio-based polymer components.
Table 2. PCM systems with bio-based polymer components.
PCMBio-Based PolymerTm [°C]Tc [°C]Heat of Melting (J/g)Heat of Freezing (J/g)Ref.
PEG 1000 Chitin nanofibers (CNFs) from crab shells of Portunus pelagicusN/A24.47N/A109.65[23]
PEG 1000Chitosan with a deacetylation degree of 80–95%407324/3175[24]
PEG 1000Microcrystalline cellulose (MCC)65.47 142.242.32137.4[25]
PEG 1000Lignin45.20 42.1643.0637.88[26]
PEG 1000Gelatin from bovine skin (Type B)33.226.461.752.6[27]
PEG 2000Spongy-like porous carbon derived from eggplants61.55 149.0019.90138.14[28]
PEG 2000PLA64.54
174.15
37.51
68.73
189.80
16.3
96.43
11.32
[29]
PEG 4000Dried pomelo peel foam 62.7 158.141.5156.1[30]
PEG 4000Enzymatic lignin6141168.7165.12[31]
PEG 4000Cellulose nanofibrils (CNF) never dried and partially delignified birch pulp fibers
ACNF: The CNF was modified by heterogeneous acetylation
LCNF: lignin-containing nanofibril suspension
58.4
58.2
N/A115.1
142.6
109.7
137.2
[32]
PEG 6000Gum tragacanth55.3106.8 39.2111.3[33]
PEG 6000Ca2+- Crosslinked SAN/A156.8N/A150.3[34]
PEG 6000PGI 58.3586.93 14.1583.65[35]
PEG 2000 to PEG 10,000Cotton composite yarn (PPCCY) was fabricated by impregnating
PEG2000–10,000 into CNT/cotton yarn (CCY) and coating electrospun PAN on its surface

Composite PPCCY with various PEG types:
PEG 2000
PEG 4000
PEG 6000
PEG 10,000
PEG 2000–56.5
PEG 4000–65.3
PEG 6000–62.9
PEG 10,000–66.3
PEG 2000–135.3
PEG 4000–136.0
PEG 6000–150.8
PEG 10,000–126.0
PEG 2000–29.8
PEG 4000–35.8
PEG 6000–35.6
PEG 10,000–39.5
PEG 2000–128.8
PEG 4000–128.4
PEG 6000–150.9
PEG 10,000–125.3
[36]
PEG 4000 to PEG 10,000To produce a PDA layer on the surface of the freeze-dried radish, the self-polymerization of dopamine was initiated 64.77161.5244.32158.41[37]
PEG 10,000CAC62.3 60.637.8 58.7[38]
Isocyanate terminated PEG 6000Pomelo peel flour (PPF) with an average particle size of about 55 μm62.5182.336.1173.5[39]
PEOCellulose (microcrystalline powder 20 mm, degree of polymerization: 229)
CAC (40% substitution acetyl groups)
CMC (60% substitution, powder < 400 mm)
CET (Cellulose ether; 30% substitution, powder 350 mm)
PEO/CEL
PEO/CMC
PEO/CAC
PEO/CET
PEO/CEL–63.4
PEO/CMC–58.4
PEO/CAC–62.7
PEO/CET–64.7
PEO/CEL–134.7
PEO/CMC–140.2
PEO/CAC–156.3
PEO/CET–156.8
PEO/CEL–32.5
PEO/CMC–36.6
PEO/CAC–27.6
PEO/CET–37.2
PEO/CEL–127.3
PEO/CMC–138.0
PEO/CAC–152.8
PEO/CET–153.3
[40]
PEOPotato starch67.396.933.594.6[7]
PEO with a viscosity average molecular weight (Mv) of 1 × 105 g/molPLLA 60.9558.7946.3157.59[41]
1-DodecanolSA25.69 149.15 ± 314.65 147.79 ± 3[42]
DodecaneCS (low MW with a degree of deacetylation of 75–85%), to obtain the biochar papermill sludge, was used as biomass−9.9N/A91N/A[43]
1-Dodecanol from coconut extract (70 wt.%)Bio-based transparent wood: delignified and succinylated birch wood25.2 ± 1 141 ± 519.3 ± 0.6 128 ± 7[44]
TD and MAHydroxylpropyl methyl cellulose (HPMC).34.61206.4531.09204.59[45]
Hexadecane, butyl stearate, caprylic acid Poly(D-galacturonic acid methyl ester)H5—23.99
B5—21.60
C5—18.50
H5—184.89
B5—116.03
C5—118.04
H5—1.49
B5—10.34
C5—−19.85
H5—185.89
B5—118.26
C5—114.42
[46]
n-Nonadecane Alginic acid sodium salt32.1028.7181.6718.67[6]
EicosaneBiochars obtained by the pyrolysis of softwood and wheat straw at 550 °C and 700 °CSoftwood biochar/eicosane–37.0
Wheat straw biochar/eicosane—37.0
N/ASoftwood biochar/eicosane–53.4
Wheat straw biochar/eicosane–75.0
N/A[47]
n-EicosaneNatural rubber-latex 35–4520–35encap-C20 prepared at polymer weight contents of 20 wt.%: 204.5 ± 8.5
and encap-C20 prepared at polymer weight contents of 9 wt.%
239.8 ± 3.8
encap-C20 prepared at polymer weight contents of 20 wt.%:208.2 ± 9.4 and encap-C20 prepared at polymer weight contents of 9 wt.%
251.4 ± 6.2 J g_1
[48]
n-Octacosaneκ-Carrageenan63.6157.751.1249.3[49]
n-OctadecaneAqueous silk solution 22.95 ± 60.734.63 ± 60.3834.98 ± 61.4037.58 ± 62.72[50]
1-Octadecanol,
1-Eicosanol,
1-Docosanol
Acrylated soybean oil (ASO)
F2—(ASO/1- Octadecanol)
F3—(ASO/1- Eicosanol)
F2—(ASO/1- Docosanol)

F2—60.46
F3—66.67
F2—70.07

F2—30.06
F3—45.58
F2—67.51

F2—49.39
F3—56.18
F2—60.82

F2—−17.56
F3—−41.01
F2—−70.07
[51]
Capric acidSilkworm cocoons30.6123.4 26.7122.6[52]
Myristic acidOrange peel biochar55.3767.204965.14[53]
Vegetable-derived PABio-based PLA62.3N/A59.9N/A[54]
CA and LASA16.54 35.1811.84N/A[55]
Sugarcane wax/Al2O3 compositeGelatin, gum Arabic68.0061.0059.6644.45[56]
Paraffin wax type HeptacosaneWaste chicken feathers77.58 N/A156.56N/A[57]
BeeswaxCoffee grounds collected after brewing in an automatic coffee machine50.1121.0834.23129.36[58]
Sodium sulfate decahydrate-Na2SO4·10H2OSA32.6N/A37.1N/A[59]
Lignin-g-PCL copolymersAlkaline lignin51.33N/A61.16N/A[60]
Erythritol (ET)-grafted-PLA Meso-erythritol, PLA 163.8170.7N/AN/A[61]
Microencapsulated PCM (mixtures of branched-chain hydrocarbons encapsulated in melamine formaldehyde shell)Microcrystalline cellulose 20 μmN/A130.34N/AN/A[62]
Microcapsule PCMs (paraffins encapsulated in melamine shells)PCM microcapsules on wool fabricN/Aenthalpy after 10 washing cycles—2.92N/AN/A[63]
Microencapsulated PCMs (paraffin in PMMA shells)Bacterial cellulose (BC)23.756.7322.586.58[64]
Microencapsulated PCM (mixtures of branched-chain hy-drocarbons encapsu-lated in melamine formaldehyde shell)Pullulan29.36104.8516.02/18.82103.58[65]
Table 3. Temperatures and heat of melting of PEG and their mixtures. Reprinted from [73] with permission from Wiley.
Table 3. Temperatures and heat of melting of PEG and their mixtures. Reprinted from [73] with permission from Wiley.
SampleHeating CycleMelting Point [°C]ΔH [J/g]
PEG 10001
2
44.2
40.0
165.3
168.6
PEG 34001
2 (2 peaks)
63.0
56.4/61.9
191.7
171.6
PEG 10,0001
2
66.6
66.2
188.7
180.6
PEG 20,0001
2
68.5
67.7
177.4
165.0
PEG 35,0001
2
71.1
68.3
199.7
183.4
PEG 3400/PEG 10,0001
2 (2 peaks)
66.4
55.6
64.3
169.2
173.5
PEG 1000/PEG 10,0001
2 (2 peaks)
36.5
60.8
35.3
161.3
156.8

PEG 1000/PEG 34001 (2 peaks)

2 (3 peaks)
33.5
60.0
34.3
51.3
59.1
199.7

183.4
Table 4. Temperature, melting, and freezing heat of the tested polyethers determined by the DSC method (heating and cooling rate 10 °C/min). Reprinted from [75] with permission from Wiley.
Table 4. Temperature, melting, and freezing heat of the tested polyethers determined by the DSC method (heating and cooling rate 10 °C/min). Reprinted from [75] with permission from Wiley.
PolyetherMeltingFreezing
Melting Range [°C]Tmax [°C]Heat of Fusion [J/g]Freezing Range [°C]Tmax [°C]Heat of Freezeing [J/g]
PEG 340051.9 ÷ 65.263.4166.830.4 ÷ 39.036.7158.3
PPO 4000−11.5 ÷ −6.0−9.20.6---
PTHF 290023.9 ÷ 33.729.791.0−3.8 ÷ 5.70.285.4
Table 5. The advantages and disadvantages of the encapsulation techniques. Reprinted from [76] with permission from Elsevier.
Table 5. The advantages and disadvantages of the encapsulation techniques. Reprinted from [76] with permission from Elsevier.
AdvantagesDisadvantages
Protection from light, heat,
moisture, and high oxygen
concentration that might lead
to the decomposition of PCMs
Micro and nanoencapsulation are complicated to conduct, and macro encapsulation has a lower structure stability and fracture resistance
Prevent evaporation of volatile compounds that harm the environment and users’ healthNeeds a standard selection
procedure of PCMs and shell
materials
Cover unpleasant odours
Control the speeds of internal
materials’ release
Low thermal conductivity due to the polymer used as the shell
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

Pielichowska, K.; Nowicka-Dunal, K.; Pielichowski, K. Bio-Based Polymers for Environmentally Friendly Phase Change Materials. Polymers 2024, 16, 328. https://doi.org/10.3390/polym16030328

AMA Style

Pielichowska K, Nowicka-Dunal K, Pielichowski K. Bio-Based Polymers for Environmentally Friendly Phase Change Materials. Polymers. 2024; 16(3):328. https://doi.org/10.3390/polym16030328

Chicago/Turabian Style

Pielichowska, Kinga, Katarzyna Nowicka-Dunal, and Krzysztof Pielichowski. 2024. "Bio-Based Polymers for Environmentally Friendly Phase Change Materials" Polymers 16, no. 3: 328. https://doi.org/10.3390/polym16030328

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