Polymers 14 00620
Polymers 14 00620
1 Department of Science and Methods for Engineering, University of Modena and Reggio Emilia, Via
Amendola 2, 42122 Reggio Emilia, Italy; [email protected] (S.B.); [email protected] (S.M.);
[email protected] (B.R.); [email protected] (M.M.)
2 Interdepartmental Research Center for Industrial Research and Technology Transfer in the Field of
Integrated Technologies for Sustainable Research, Efficient Energy Conversion, Energy Efficiency of
Buildings, Lighting and Home Automation, EN&TECH, University of Modena and Reggio Emilia,
Via Amendola 2, 42122 Reggio Emilia, Italy
3 TekneHub Laboratory, University of Ferrara, Via Saragat 13, 44122 Ferrara, Italy;
5 Interdepartmental Center for Applied Research and Services in Advanced Mechanics and Motoring,
INTERMECH-Mo.Re., University of Modena and Reggio Emilia, Via P. Vivarelli 10/1, 41125 Modena, Italy
* Correspondence: [email protected]
Citation: Barbi, S.; Barbieri, F.; Abstract: The building sector is responsible for a third of the global energy consumption and a
Marinelli, S.; Rimini, B.; Merchiori, quarter of greenhouse gas emissions. Phase change materials (PCMs) have shown high potential for
S.; Bottarelli, M.; Montorsi, M. Phase latent thermal energy storage (LTES) through their integration in building materials, with the aim
Change Material Evolution in
of enhancing the efficient use of energy. Although research on PCMs began decades ago, this
Thermal Energy Storage Systems for
technology is still far from being widespread. This work analyses the main contributions to the
the Building Sector, with a Focus on
employment of PCMs in the building sector, to better understand the motivations behind the
Ground-Coupled Heat Pumps.
restricted employment of PCM-based LTES technologies. The main research and review studies are
Polymers 2022, 14, 620. https://
doi.org/10.3390/polym14030620
critically discussed, focusing on: strategies used to regulate indoor thermal conditions, the variation
of mechanical properties in PCMs-based mortars and cements, and applications with ground-
Academic Editors: Mariaenrica
coupled heat pumps. The employment of materials obtained from wastes and natural sources was
Frigione, Giovanni Dal Poggetto and
also taken in account as a possible key to developing composite materials with good performance
Cristina Leonelli
and sustainability at the same time. As a result, the integration of PCMs in LTES is still in its early
Received: 26 January 2022 stages, but reveals high potential for employment in the building sector, thanks to the continuous
Accepted: 3 February 2022 design improvement and optimization driven by high-performance materials and a new way of
Published: 5 February 2022 coupling with tailored envelopes.
Publisher’s Note: MDPI stays
neutral with regard to jurisdictional Keywords: phase change materials; latent thermal energy storage; sustainable buildings;
claims in published maps and ground-coupled heat pumps; energy reduction; materials design; eco-friendly materials;
institutional affiliations. sustainable materials; green economy
greenhouse gas emissions [1]. Despite a flattening of CO2 emissions between 2013 and
2016, building-related emissions began to grow back in recent years, leading to a total
increase equal to 5% from 2010 to 2019. Fossil fuel use has also grown, at an average rate
of 0.7%/year, in the same period [2]. The reduction of greenhouse gas emissions and the
rational use of energy have become a global widespread endeavor, especially in the
European Union (EU), where a relevant need for energy saving in buildings has been
recognized and put into practice. In fact, the more recent EU environmental policies for
the building sector have promoted the spread of renewable energy technologies [3,4]. The
need to enhance sustainable energy employment in this field could be satisfied through
different approaches and technologies.
One of these technologies is thermal energy storage (TES), which allows the storage
of heat or cool energy in well-tailored materials during low demand periods to release
them when needed. The storage mechanism in TES architectures is usually carried out
using the sensible or latent heat of specific materials, taking advantage of the specific heat
and temperature change with the former, and of the phase-change latent heat performed
at a given temperature with the latter [5,6]. Sensible heat TES (SHTES) architectures
generally use cheap storage materials, such as water, bricks, or rocks, and for this reason,
they are established and widespread technologies. Latent heat TES (LHTES) systems, by
contrast, are based on phase change materials (PCMs) and offer the advantages of a fairly
constant working temperature and the enhanced energy density of their storage material,
which allows the storing of 5–14 times more energy than SHTES in the same volume,
therefore reducing the size of the storage system [7,8]. In fact, PCMs can be organic,
inorganic, and eutectic compounds characterized by high values of phase-change latent
heat. The main classes of PCM are shown in Figure 1.
The majority of organic PCMs are composed of paraffin waxes (CH3(CH2)nCH3), with
melting temperatures between 20 °C and 70 °C, thereby fulfilling the needs of LHTES
architectures in the buildings sector in a wide range of latitudes [6]. Fatty acids
(CH3(CH2)nCOOH) constitute the second main group of organic PCMs, with melting
temperatures between 17 °C and 64 °C. Other organic substances generally exploited as
PCMs are esters, alcohols, and glycols [6,10]. Among inorganic PCMs, hydrated salts are
the main category. With respect to organic compounds, they have a wider range of
Polymers 2022, 14, 620 3 of 33
melting points, reaching temperatures up to 140 °C [11,12]. Molten salts have also been
considered, but their high melting temperatures (>250 °C) make them unattractive for the
building sector [13]. Finally, eutectics can also be considered, as they are made of mixtures
of two or more pure PCMs, which can be organic–organic, inorganic–inorganic, or
inorganic–organic. They show well-defined melting points, but their properties and
performances must be further investigated as they are the result of very specific
formulations that could be difficult to control for industrial purposes and when handling
large volumes of material [9,14].
Every category of PCM, taking into account the solid-liquid latent heat, presents
several advantages but also drawbacks, as listed in Table 1, which limit their practical and
commercial applications in the building sector [15]. There are also some limitations
common to all categories, namely high costs, low thermal conductivity, and long-term
stability, which make SHTES even more widespread and profitable than LHTES [16].
Among phase changes, solid–liquid is the most employed in practice. In fact, even if the
latent heat of the liquid–gas transition is generally higher, the relevant associated volume
change must be considered, as it can lead to many severe critical issues in the design of
liquid–gas-based TES, such as partially flexible PCM containment or safety devices due
to slight temperature changes [6]. On the other hand, PCMs based on solid–solid
transition take advantage of the passage between a crystalline and an amorphous solid
phase. The lack of a liquid or gas phase prevents leakage problems, but this PCM category
shows a lower phase change energy compared to PCMs based on solid–liquid transition.
Moreover, their phase change temperatures are usually higher than those required for
building applications [9,17].
Table 1. Principal advantages and disadvantages of organic, inorganic, and eutectic PCMs [15].
Since the early studies from 1990 onwards, the use of LHTES technology was
considered in the building sector to enhance energy management and consumption,
dividing the final PCM applications (for heating or cooling purposes) into active and
passive [6]. The significant difference lies in the fact that in active applications, the thermal
energy is available only when necessary, using mainly on/off auxiliaries such as heating,
ventilation, and air conditioning systems (HVAC). Passive applications, instead, aim to
include PCMs directly into building elements, using only the surrounding environment
as their energy source, allowing automatic and continuous energy storage or release and
providing thermal comfort with minimum or no use of HVAC systems [18,19].
Active applications include the use of TES coupled with conventional HVAC systems
(e.g., floor or ceiling heating plants) or with heat pumps to optimize the electric energy
demand from high-peak to off-peak periods [20]. Other active applications consider solar
energy systems and solar air collectors integrated into building walls, as well as the use
of ventilation mechanisms with free cooling, that take advantage of low night
temperatures to store coldness in building materials and release it during the day.
Passive technologies include standard solar walls, so-called Trombe walls, and the
inclusion of PCMs in construction materials such as concrete, mortars, and insulation
materials, or in building elements, such as windows, shutters, wallboards, and bricks [20].
Polymers 2022, 14, 620 4 of 33
An important aspect in all the applications is that the employed PCM must be tai-
lored for a specific use, considering its nature (organic or inorganic), its percentage in the
formulation, and, especially, its precise melting temperature according to climatic condi-
tions, building design, and thermal comfort requirements. As PCMs can be fully inte-
grated with the TES architecture, the environmental sustainability of all the system must
also be taken into consideration. Several environmental analyses based on the life cycle
assessment (LCA) methodology have shown that the environmental impact resulting
from the production, installation, and disposal of PCMs is largely recovered from the en-
vironmental benefit obtained thanks to energy savings (from 15% to 35% of energy saved
based on climatic conditions) [21]. However, other studies have shown that the use of
PCMs as paraffins or salt hydrates, while reducing energy consumption, does not signifi-
cantly reduce the overall impact compared to the use of a conventional insulation mate-
rial, such as polyurethane, suggesting that the production and the disposal phases of PCM
life cycles need to be further investigated [22]. For example, paraffin production is mainly
based on fossil oils, although materials derived from renewable sources have also recently
been developed [23]. The production process is very similar for all fossil-based materials,
regardless of the length of the carbon chain that determines the melting point. The envi-
ronmental impact can therefore be considered identical for the individual materials [24].
Thereafter, more recent studies focus on more sustainable organic materials not based on
fossil fuels. For example, the biodegradability of a plant-based PCM was analyzed and
compared to paraffins and salt hydrates, demonstrating the total decomposition of the
plant-based material after 4 months [25].
Despite the considerable number of studies on PCM integration in LHTES systems,
elements, or materials, their application in the building sector is far from being wide-
spread. The aim of this work is to analyze the developments of PCM applications in the
building sector, since their early stages, through an analysis of research and review pa-
pers. In the first period (2000–2010), described in the Section 2, the main focus was on the
improvement of PCMs’ general properties. Next, a focus on the more recent and peculiar
applications in the building sector is presented, according to the materials’ evolution in
recent years. In fact, during 2010–2021, research in the building field was mainly driven
by PCM application, as very different effects can be exploited depending on the relative
placement of PCMs in a building or in specific climatic areas. Finally, research concerning
the coupling of PCMs to ground-coupled heat pump (GCHP) technology is discussed. In
fact, this technology, by employing the soil as a form of underground thermal energy stor-
age, is one of the most promising and innovative LHTES architectures, as it is not influ-
enced by seasonal effects.
available products (29 and 61, respectively). Nevertheless, several critical points common
to the majority of PCMs arise, such as long-term stability, low thermal conductivity, den-
sity variation over the transition phase, corrosion, and subcooling or phase segregation.
These drawbacks are particularly relevant for inorganic PCMs, including eutectic mix-
tures, such as LiF–CaF2, and limit their employment despite the higher value of their latent
heat. In particular, LiF–CaF2 eutectic mixtures have shown a linear coefficient of expan-
sion near to 30 × 10−6 K−1 and relevant changes over time (after 9 months) in density (−2%)
and porosity (+2.1%) [31]. In addition, the numerical simulation and modeling of the heat
transfer during the phase change, called the moving boundary problem, was also consid-
ered as a critical point to study, over several work cycles, to validate numerical simula-
tion, due to a lack of experimental studies [14,26,32].
As a possible strategy to overcome PCMs’ intrinsic low thermal conductivity, highly
tailored encapsulation methods were investigated, as a promising way to increase PCMs’
heat transfer while avoiding large phase separations [28]. In particular, macro-encapsula-
tion was analyzed as a viable and economical way of making PCM employment more
efficient, giving them a self-supporting structure designed for their intended application.
Examples of macro-encapsulation strategies are shown in Figure 2. They include metal
tubes (Figure 2a), metal spheres (Figure 2b) that could be arranged to form a packed bed,
PCM spheres (Figure 2c), aluminum pouches (Figure 2e), and rectangular flat panels in
polyvinyl chloride (PVC) or aluminum (Figure 2d,f).
These different tested geometries show different peculiarities. With flat panels and
pouches, it is possible to achieve more surface area per unit of storage volume, having less
weight and volume with respect to bulk PCMs. Subsequently, the solidification time re-
sults reduced to 4 h and the melting time reduced to 6 h, making it acceptable for free
cooling by employing a low PCM thickness [34,35]. Nevertheless, the panel geometry was
Polymers 2022, 14, 620 6 of 33
found to be superior to pouches as, under the same environmental conditions and by em-
ploying a commercial organic PCM, the first showed higher cooling power (+12.5%) and
shorter melting times (−45%) [36]. Cylindrical PCM pipes are easier to fabricate, having
the same heat transfer characteristics with respect to flat panels and a lower heat loss rate
[33,37,38]. In fact, in this geometry, heat transfer takes place in the axial and radial direc-
tions, with an increased area of convective heat transfer. Concerning spherical geometry,
PCMs or metals balls have larger surface areas per unit of volume compared to cylindrical
geometry and heat transfer can be tailored by selecting the ball’s diameter. For example,
balls up to 3 mm in diameter were tested, demonstrating that the solidification time is
about 1.5 h, which is suitable for free-cooling applications [36]. In particular, a LHTES
configuration based on PCM spherical geometry encapsulation (Figure 3) has been inves-
tigated as a possible solution to increase the heat transfer rate between the PCM and the
heat transfer fluid (HTF) while avoiding direct contact between them[28]. However, the
main problem of the packed-bed configuration concerns the complex modeling of both
the heat transfer process between PCM and HTF considering the container and the PCM
phase change inside the container itself [28].
When sphere geometry concerns restrained diameters (<5 μm), the micro-encapsula-
tion definition is employed. Micro-encapsulation architecture, thanks to the larger heat
transfer area between micro-encapsulated particles, ensures an even greater thermal con-
ductivity with respect to sphere geometry encapsulation (or macro-encapsulation). How-
ever, a further encapsulation was needed, in these years, to collect all the micro-spheres,
consequently reducing the micro-encapsulation’s beneficial effects on the final thermal
conductivity dramatically [28]. In fact, further collection of the micro-capsules forces the
majority of the heat transfer to occur through conduction, significantly limiting the overall
heat transfer rate. In this context, for macro-encapsulated PCMs, the PCM-HTF heat trans-
fer process modeling, as well uncertain long-term stability, requires more investigation
due to micro-encapsulation architecture [39]. In addition, cost must also be considered,
and at the early stage of the research this was among the most expensive storage methods
for PCMs [28].
To overcome micro-encapsulation issues, more complex geometries, such as the use
of fins or extended surfaces in the PCM, to provide additional heat transfer surface, were
studied in order to enhance thermal conductivity [29]. Nevertheless, it has been demon-
strated that their employment must be finely tuned with respect to their final application,
as the beneficial effect of fins decreases with increases in the heat transfer coefficient. For
Polymers 2022, 14, 620 7 of 33
this reason, in most cases, fins must be included at the side of the PCM container and their
use must be regulated if melting or solidification phases are to be enhanced [29]. In addi-
tion, fins can vary widely in material, number, and thickness, as these parameters are cru-
cial to promote both conduction and convection processes during PCM phase change.
From several studies, it was demonstrated that heat transfer during PCM melting is
mainly ruled by natural convection, therefore, in this case, fins could be helpful to further
promote the natural convection process [29]. In fact, generally, the same material as the
PCM’s envelope is employed. Rectangular fins are the most frequently used due to their
very simple geometry and low-cost manufacturing, but ring-shaped fins were also em-
ployed in configurations in which the PCM is placed around HTF tubes, as shown in Fig-
ure 4. Numerical models that compared the latter solution with a system without fins,
demonstrated that the great majority of heat was conducted through the fins along the
radial direction. For different ranges of mass flow rate and inlet temperature, a consider-
able amount of increase in the energy stored was observed due to the presence of fins,
increasing with the fins’ number, up to + 63% with 19 fins [29,40].
The impregnation of porous structures with PCMs was investigated as another viable
way to increase their thermal conductivity. However, the final result depends not only on
the materials’ thermal conductivity itself, but also on their mean pore size [29]. In fact,
several numerical studies demonstrated a substantial decrease in the response time of the
porous metal matrix-PCM system; consequently, a high porosity (up to 85%), for the ob-
tainment of the highest possible performance, must be promoted. However, this solution
raise concerns related to the final disposal of the porous structure, as the PCMs’ presence
inside the porosity must be considered and evaluated at the end-of-life when recycling or
landfill disposal are needed. The most commonly investigated materials were metals, e.g.,
aluminum or copper, and carbon-based materials, such as graphite, which presents a nat-
ural porous structure, or expanded graphite (EG) [29,41].
Metals can also be added to the PCM in the form of high-conductivity particles, or as
metal structures (e.g., metal rings or screens). However, particle mass fraction must be
carefully designed to enhance thermal conductivity, without lowering the thermal storage
capacity of the PCM. In fact, metal structures usually occupy much more volume (20%) in
the PCM with respect to other technologies, such as fins (7%). Copper, aluminum, and
silver are the most studied metals for applications with particles, due to their high cost
and enhanced thermal conductivity, while steel is preferred for metal structures or ex-
tended geometries. With metals, only paraffin-based PCMs can be used, because of the
corrosion phenomenon due to salt hydrates [26,29]. Among carbon-based materials, char-
acterized by high thermal conductivity and low density, carbon fibers have been consid-
ered, as they can be added in the PCM in form of fibers, brushes, or cloths. This material
allows a good enhancement of thermal conductivity with a volume fraction of approxi-
mately around 1% and a reduction in solidification time of around 23%, but also presents
high costs due to the need for a very uniform distribution of fibers in the PCM [29].
Polymers 2022, 14, 620 8 of 33
In these years, to overcome these challenges in PCM design and optimization, the
mathematical and numerical modeling of phase change and heat transfer phenomena that
take place in LHTES systems were approached first. Two main models and related cate-
gories were developed during the 2000–2010 period: those based on the first law of ther-
modynamics, which aimed to quantify the amount of stored energy; and those based on
the second law, which aimed to quantify the usefulness of stored energy. Consequently,
the two models were concern with energy and exergy analysis, respectively. Most of the
studies in this period focus only on energy analysis, while exergy-based evaluation is less
frequently considered, despite its value and complementarity with the first. This trend is
due to the fact that most energy-based models are validated through experimental set-
ups, while the acceptability of exergy-based models suffers from a lack of evidence against
experimental data [27,41]. Historically, simplified analytical approaches were proposed
with the so-called Stefan problem and Neumann method, as clearly reported by GSH Lock
[42]. However, to comprehend the behavior of these phenomena, numerical codes are
needed, as they formally implement and solve energy and heat transfer equations, based
on different finite or volume element numerical approximations.
The research on PCM applications in the building sector was at its very early stage
during the period of 2000–2010. Trombe wall and building block applications were inves-
tigated with both organic PCMs, such as paraffin wax and stearic acid, and inorganic salt
hydrates. In this context, several studies demonstrated that for a given amount of heat
storage, the addition of PCM (~20% of the weight being paraffin) requires less Trombe
wall volume (−20%) and weight. At the same time, concrete and wallboard functionaliza-
tion (direct mixture), mainly with mainly organic PCMs, such as paraffin waxes or fatty
acids, were considered [6,30]. Despite the promising results of these studies in terms of
good heat transfers, some critical issues must be highlighted. The most important of all
was the small number of suitable commercial materials, which must have both a transition
temperature falling in the human comfort range (20–32 °C) and a reasonable cost of em-
ployment in amounts high enough to be mixed with concrete [6,9]. In addition, to enhance
the free cooling phenomenon, the combination of at least two PCMs was suggested to
allow year-round thermal management [33]. In fact, free cooling takes advantage of low
night temperatures to store coldness in building materials enriched with PCMs, or in
LHTES units, and release it for indoor cooling during the daytime and, for this reason, is
strongly influenced by seasonal variation. Thereafter, it is almost impossible to choose
only one PCM as it would melt/solidify only for a part of the year.
Table 2. PCM classification according to different application types and temperature range in the
building sector up to 2010, according to Cabeza et al. [43].
Hot-Water High-Temperature
Cooling Comfort Applications
Applications Applications
Temperature range −30/+21 °C +22/+28 °C +29/+60 °C +61/+120 °C
Total PCM number 45 34 103 62
Organic (paraffins, fatty acids,
23 (51.1%) 22 (64.7%) 50 (48.5%) 28 (45.2%)
organic mixtures) (%)
Inorganic (salt, salt hydrates,
6 (13.3%) 7 (20.6%) 45 (43.7%) 19 (30.6%)
metals, inorganic mixtures) (%)
Eutectics (%) 14 (31.1%) 5 (14.7%) 8 (7.8%) 15 (24.2%)
As shown in Table 3, the number of commercial PCMs available in 2010 was limited
to 88 products, which is quite low if compared with the 244 total PCMs in Table 2, as
confirmed by Zhou et al. [44]. In addition, it can be noted that at this stage, no commercial
PCM had been developed for hot-water applications; nevertheless, the great majority of
the total number of PCMs were devoted to this application (Table 2). A possible reason
for this trend is that commercial products had been developed for more recent and ad-
vanced applications at that time, while hot-water applications were the most consolidated.
Later studies reported a strong acceleration of the development of new PCMs dedicated
to the building sector. A comparison between commercial products is represented in Fig-
ure 5, which shows that in the same temperature range (15–33 °C), the number of com-
mercial PCMs grew from 20 to 90 in 4 years, from 2011 to 2015. The growth concerns all
PCM categories, but organic substances show the highest increase during these years with
great variability in the technology used, e.g., tubes or mats filled with PCMs, gypsum
boards, floor tiles, windows, heat exchangers, or ventilation systems with PCMs [43,45].
Table 3. Classification of commercial PCMs according to different application types and tempera-
ture ranges in the building sector up to 2010, according to Cabeza et al. [43].
Higher-Temperature
Cooling Comfort Applications
Applications
Temperature range −33/+21 °C +22/+28 °C ≥29 °C
Total PCM number 24 13 51
Organic (paraffins)
2 (8.3%) 5 (38.5%) 29 (56.9%)
(%)
Inorganic (salt
solutions, salt 22 (91.7%) 8 (61.5%) 20 (39.2%)
hydrates) (%)
Unclassified (%) 0 (0.0%) 0 (0.0%) 2 (3.9%)
Polymers 2022, 14, 620 10 of 33
Figure 5. Comparison between 2011 and 2015 of commercial PCMs useful for building applications,
in the 15–33 °C range [43,45].
Some of the first studies in this period (2011–2012) still focused more on the issues
concerning PCMs’ intrinsic properties than on the problems facing their application in
building materials [43,44]. The main critical points that arose in this period were: the need
to further enhance the heat transfer due to PCMs’ low thermal conductivity (from 0.2
W/m⋅K for paraffins to 0.5 W/m⋅K for hydrated salts); the long-term stability of PCMs with
an increasing number of thermal cycles; the compatibility between the PCM and its con-
tainer, such as corrosion between salt hydrates and metal and the dimensional stability of
plastics containing organic PCMs; and the phase segregation or subcooling of salt hy-
drates [43,44]. In fact, in this period, investigations tried to exploit salt hydrates’ peculiar-
ities due to their intrinsic higher phase change enthalpy, by reducing their possible draw-
backs. Later studies in this period show that these problems were not completely solved
for specific categories of PCM, even in the following years.
In fact, the long-term stability of paraffins was assessed, as all the materials tested
showed almost no changes in their properties after a number of thermal cycles, between
300 and 5000 [46]. Different studies demonstrated that the stability of fatty acids depends
instead on their purity degree, as compounds with a purity degree between 90% and 95%
(e.g., stearic acid, palmitic acid, myristic acid, or lauric acid), showed a latent heat drop
after 900–1200 thermal cycles [46]. Phase separation and the supercooling of salt hydrates
were improved, respectively, by employing thickening agents (e.g., attapulgite clay, pol-
yvinyl alcohol, silica gel) and nucleating agents (e.g., Borax, sodium pyrophosphate dec-
ahydrate), subsequently leading to improved long-term stability (e.g., CaCl2 6H2O can
reach 1000 working cycles without phase separation by addition of NaCl and H2O) but
with a more complex PCM formulation to keep under control during their fabrication
[45,46].
With respect to the previous period (2000–2010), in these years (2011–2015), more
specific ways were found to incorporate PCMs into building elements. According to Zhou
et al. [44], five main methods can be defined, as follows:
- Direct incorporation: the PCM and the building material are physically mixed at the
solid state.
- Immersion: a building element, e.g., bricks or concrete, is dipped into a liquid PCM,
which enters into the open pores thanks to capillary forces.
- Macro-encapsulation: the PCM is encapsulated inside a container with a variable
shape.
Polymers 2022, 14, 620 11 of 33
also plays a very important role, driving microencapsulated PCMs’ morphologies, shell
mechanical strength, heat capacities and thermal stabilities. Among the organic materials
that are easiest to manipulate for microencapsulation fabrication, the most frequently em-
ployed are: polystyrene, polyurea, polymethyl methacrylate, arabic gum, amino resin,
urea formaldehyde resin, melamine, and formaldehyde resin. Nevertheless, their employ-
ment should be restricted due to their flammability, toxicity, and low heat conductivity.
By contrast, the inorganic materials employed are mainly silica-based due to their desira-
ble properties, such as their chemical and thermal stability, flame retardancy, high storage
capacity and good compatibility with building materials [48]. Due to this complexity in
the period investigated, a new category of PCMs arose, named composite PCMs. Moreo-
ver, direct solar gain and the ventilation rate should be considered together as environ-
mental temperature fluctuation, as they can affect the effective temperature measured on
the wallboard, leading to over or underestimation of the PCMs’ performance. Finally, it
must be noted that, in several studies, the amount of the building element (e.g., a wall)
covered with PCM, or with PCM added to it, is not always specified, and this leads to an
increased difficulty in quantifying the real performance of the PCM, which also makes it
more challenging to compare between different studies [10,44,48,49,51,52].
The integration of PCMs into floors and ceilings is usually intended to charge the
materials with heat or coolness during the night and release them during the day, shifting
the peak of electricity consumption to the off-peak period and allowing the use of cheaper
energy. Wallboards have also been used in ceilings, coupled with active systems to en-
hance indoor thermal regulation. Metal containers (typically iron or steel) filled with salt
hydrates or micro-encapsulated paraffin mixed with gypsum or rock wool were placed
on the ceiling and coupled with active systems, as a ventilation system or a capillary sys-
tem of water tubes. The results demonstrated that these systems are able to store coolness
during the night and release it during the day, reducing the maximum peak temperature
by up to 2 °C [48]. A similar principle is used in underfloor heating or cooling systems,
which could be coupled with PCM plates to store heat or coolness [10,44,48].
Different studies investigated the addition of PCMs into concrete matrices, as they
are characterized by a high thermal mass, which allows the storage of thermal energy.
Furthermore, for this application, usually, only organic PCMs are considered, e.g., paraf-
fin, butyl stearate, dodecanol, and tetradecanol [53]. One of the main critical points inves-
tigated, in greater depth than in previous years, was how to promote the addition of PCMs
into the concrete matrix, avoiding an excessively strong decrease in mechanical proper-
ties. In fact, although the addition of micro-encapsulated PCMs seems to consistently re-
sult in the worsening of mechanical properties, the final product is still appropriate for
several building purposes. It has been demonstrated that a compressive strength of over
25 MPa and a tensile splitting strength of over 6 MPa, without any variation after 6
months, can be reached by adding a 5 wt% of commercial PCM called Micronal from BASF
[44,53]. On the other hand, other studies report a 13% decrease in compressive strength
for each 1% of micro-encapsulated paraffin added to the matrix, or a decrease equal to
32% for a 5% addition of micro-encapsulated PCM [48,53]. The difference among these
results must be founded in the fact that, during these years, specifically tailored commer-
cial materials have been developed in order to limit the concrete’s property loss. The loss
of workability and the higher costs connected to the addition of PCM to concrete are ad-
ditional problems, and the limited percentage of material that could be added could result
in a low increase in the total heat storage capacity for concrete [49]. Furthermore, micro-
encapsulated PCM addition between 1% and 5% grants an increase in concrete-specific
heat capacity; it is also the cause of lower thermal conductivity. Experiments performed
on concrete cubicles (Figure 6) with 5% paraffin-based commercial PCM demonstrated
that summer peak temperatures could be lowered by up to 3 °C with PCM addition when
compared with concrete only [51,53]. On the other hand, it is not clear whether this lower
peak was due to the higher thermal capacity or to the lower thermal conductivity of the
material, although this performance can be repeated every day only if the PCM is allowed
Polymers 2022, 14, 620 13 of 33
to completely solidify during the night [44,48,49,51,53]. In this context, PCM leakage out
of micro-capsules must also be considered as a potentially critical issue, not only for over-
all technical performance but also for environmental safety [51].
of salt hydrates or ester as PCM led to impact reductions of 9% and 10.5% respectively,
compared to the case of using paraffin [55].
Table 4. PCMs studied for building applications in the period 2016–2021, according to Akeiber et
al. [19], Liu et al. [56], Singh Rathore et al. [57], and Da Cunha and De Aguiar [58].
Melting Point
Category Encapsulation Type Shell Material Reference
(°C)
RT 18 Organic 18 Macro Steel [56]
Capric acid and lauric
Fatty acids 20 Macro Stainless steel [56,58]
acid
EPDM and furnace
RT 21 Paraffin 21 Macro [19]
dust
SP 22 Inorganic 21 Macro / [19]
Hexadecane Paraffin 22 Macro Copper [56]
Micronal DS-5008X Organic 23 Micro Acrylate polymer [19,57,58]
Inertek Organic 23+27 Micro Polymer [58]
PEG 600 Polymer 21–25 Macro PVC [56–58]
Micronal DS 5001 X Organic 23–26 Micro Acrylate polymer [57,58]
Capric acid and 1- Fatty acid and
26 Macro Aluminum [19,56]
dodecanol fatty alcohol
Dodecanol Fatty alcohol 26 Micro / [57]
Capric acid and
Fatty acids 26 Micro Polystyrene [57]
myristic acid
Capric acid and
Fatty acids 26 Macro Gypsum wallboard [19,56]
palmitic acid
SP 25 Salt hydrate 26 Macro Aluminum [56,58]
Calcium chloride
Salt hydrate 25–27 Macro PVC [19,56]
hexahydrate
Macro Aluminum
RT 27 Organic 28 [56–58]
Micro /
RT 28 Organic 28 Micro / [57]
RT 28HC Organic 27–29 Macro Aluminum [19,56,58]
MG29 Paraffin 27–29 Macro Glass [19]
Micro CaCl2
Octadecane Paraffin 28–29 [57]
Micro PMMA/tio2
Micro Brookite (TiO2)
Eicosane Paraffin 30 [57,58]
Macro Galvanized steel
Capric Acid Fatty acid 30 Macro Aluminum [19,56]
Salt hydrate Salt hydrate 31 Macro Polymer [56]
RT 35 Organic 28–35 Macro Aluminum [19,56]
Tetradecanol and Fatty alcohol and
29–32 Macro PE-RT [56]
myristic acid fatty acid
MPCM37-D Paraffin 37 Micro Polymer [57]
RT 42 Organic 38–43 Macro Stainless steel [56]
Polymers 2022, 14, 620 15 of 33
According to Table 4, which reports a summary of the most studied materials accord-
ing to Akeiber et al. [19], Liu et al. [56], Singh Rathore et al. [57], and Da Cunha and De
Aguiar [58], the majority of experimental studies still consider organic PCMs. Paraffins
(e.g., hexadecane, octadecane, and eicosane) are the most investigated, followed by fatty
acids (e.g., capric acid, lauric acid, and palmitic acid). The phase change temperature of
PCMs usually varies between 18 °C and 30 °C, but there are some applications with higher
temperatures, between 37 °C and 43 °C, according to climate changes and particular
needs. The same trend was also confirmed by Song et al. [59], who list the applications of
PCMs in different parts of the building envelope, namely walls, ceilings, and floors. How-
ever, the distribution of studies between different building elements is not equal. Accord-
ing to Zhu et al. [60] and Da Cunha and De Aguiar [58], the majority of studies concern
wall applications (approximately 60% of studies), followed by ceiling and floor applica-
tions, which mainly include the introduction of PCMs into concrete, gypsum boards or
panels. Considering Table 4, macro-encapsulation techniques were still favorable with re-
spect to microencapsulation, due to the still-high cost of the latter.
An additional focus on the experimental studies performed on PCM applications be-
tween 2019 and the first half of 2021 is presented in Table 5 (passive applications) and
Table 6 (active applications). The trend presented by the studies in Table 4 is mainly con-
firmed, as paraffins were the most studied PCMs during these years, with some excep-
tions, such as tetradecanol and hexadecanol or coconut oil. All the materials studied are
organic, with a small number of exceptions that consider two salt hydrates, namely CaCl2
∗ 6H2O and MgCl ∗ 6H2O. Nearly half of the studies still consider macro-encapsulated
PCMs, while micro- or nano-encapsulated PCMs are a minority, as shape-stabilized
PCMs. The phase change temperatures range from 17 °C to 44 °C, and walls are again the
most studied building elements; however, the inclusion of PCMs in the ceiling and floor
is also studied. Most of these studies investigate the performance of building elements in
passive conditions in field tests, attempting to understand whether the inclusion of PCMs
alone could prevent or minimize the use of active systems to regulate indoor thermal con-
ditions in different climates, as well as avoiding energy consumption.
Table 5. Main experimental studies on passive applications performed on building elements con-
taining PCMs between 2019 and 2021.
Building Position in
Reference Test Scale PCM PCM Form Tmelt (°C) Remarks and Results
Material Building
The PCM decreased the cooling load, peak,
aluminum
[61] field test RT28HC macro 28 wall and average temperature in the room by
panel
0.8 °C when coupled with the radiative panel.
The PCM blind was used to reduce the
blind in a
aluminum overheating problem typical of double skin
[62] field test PX 35 micro 35 double-skin
panel facades in summer, stabilizing the internal air
façade
temperature between the two glass layers.
The PCM roof allowed a reduction of indoor
air temperature fluctuations from 7% to 15%.
SSPCM Combined with a high-reflectivity roof, it
[63] field test paraffin (paraffin + 25.5 concrete panel roof reduced the indoor air temperature
graphite) fluctuation from 8.5% to 17.0%, while the
inner surface temperature of the roof was
reduced by 2.2 °C.
The integration of PCMs reduced both indoor
and materials’ temperature. Cool roof
polyurethane materials benefit from lower phase change
[64] lab test paraffins SSPCM 25, 31 and 44 roof
membrane temperatures (25–35 °C), while common dark
membranes show a better performance with
higher temperatures (31–45 °C).
macro PCM reduced the indoor average
[65] lab test CaCl2 ∗ 6H2O (Polyvinyl 26 wallboard Wall temperature and its fluctuations, but this was
chloride) highly dependent on climate conditions.
Polymers 2022, 14, 620 16 of 33
Table 6. Main experimental studies on active applications performed on building elements contain-
ing PCMs between 2019 and 2021.
Building Position in
Reference Test Scale PCM PCM Form Tmelt (°C) Remarks and Results
Material Building
BioPCM was not efficient for cooling during
wallboard
paraffin (DuPont SSPCM summer, but it was found to be effective in
laminated with
Energain). Soy (DuPont 21.6 winter, to increase the thermal inertia of
aluminum
[74] field test and palm oil Energain) (Energain) 25 walls, ceiling buildings.
(Energain)
(BioPCM Q25 macro (BioPCM) Energain reduced the peak temperature by
pouches
M51) (BioPCM) 3–4 °C and the daily temperature
(BioPCM)
fluctuations to 1–2 °C.
The PCM increased maximum floor
[75] lab test paraffin SSPCM 17.2 resin sheet floor temperatures during heating and cooling
respectively by 5.04 °C and 1.08 °C. The
Polymers 2022, 14, 620 17 of 33
As previously suggested, in this period, one of the principal innovations was due to
the introduction of PCMs to mortars. The introduction of PCMs into concrete and mortars
was reviewed by Rao et al. [79] and Berardi and Gallardo [80]. Different kinds of mortar
were considered, but cement mortar was the most studied, as shown in Figure 7, com-
pared to other types, such as lime mortar, aerial lime, and gypsum, or geopolymer mortar.
Figure 7. Number of different combinations of PCMs and mortars investigated in the period 2016–
2021 [79].
From the analysis of these studies, paraffins were the most frequently studied of the
PCM categories with the potential to be incorporated into mortars, as they do not disturb
the hydration reactions. On the other hand, common direct incorporation methods, in-
cluding the wet mixing or immersion technique, usually imply a high risk of leakage of
the PCM in its liquid phase, or an interaction with the matrix during its lifetime, and this
problem is quite critical for organic PCM employment [58,80]. Other direct incorporation
methods tested require a supporting material, which is often cheaper than those used for
macro or micro-encapsulation, to avoid PCM leakage [79,80]. Silica fume and expanded
graphite were studied as supporting materials, as they also enhance the thermal conduc-
tivity of the PCMs. Other materials studied were expanded perlite or vermiculite,
Polymers 2022, 14, 620 18 of 33
lightweight aggregates (LWA), or rice husk aggregates, as they are chemically compatible
with mortars [79,80]. The use of paraffin and expanded graphite showed good chemical
compatibility with lime mortar, but involved too high a cost due to the micro-encapsula-
tion process [79]. LWAs (based on clay, pumice, diatomite, and expanded vermiculite or
perlite) were enriched with PCMs through the vacuum impregnation technique, which
enables the absorption of PCM up to 73 wt%, and then coated with epoxy or polyester
resin, or limestone powder, to avoid leakage [80]. However, it was demonstrated that the
use of these coatings reduced the thermal conductivity of the whole material, while the
addition of LWAs enriched with PCMs is reported to reduce the compressive strength of
mortars from 10% to 72%, depending on the LWA material and the PCM percentage
added [80]. Nevertheless, the use of expanded perlite (30 wt%) was successfully tested to
control the leakage of paraffin [80].
The use of micro- and macro-encapsulation was also studied to add PCMs into mor-
tars. For the first, commercial polymeric microencapsulated PCMs were employed, often
as a replacement of a part of fine aggregates used in concrete or mortars, such as sand.
The results showed that the amount of micro-encapsulated PCMs should not exceed 6
wt% of concrete (which corresponds to 10–12% of volume) to limit the reduction in the
workability and compressive strength of the whole material, which is usually around 40%
for the compressive strength when the PCM weight percentage is higher than 3% [80]. The
use of micro-encapsulated PCMs also reduces the thermal conductivity of the material
with respect to the concrete without any PCM. In fact, despite the higher specific surface
of micro-capsules with respect to macro-capsules, they have lower thermal conductivity
than common sand. Moreover, their addition increases porosities of the concretes, which
are filled with low-conductive air. This effect causes a drop from 25% to 50% in the ther-
mal conductivity of concrete when a 5% mass of PCM is added [58,80]. Macro-encapsu-
lated PCMs employ stainless steel to encapsulate the PCM, leading to many advantages,
such as lower leakage compared to micro-encapsulated PCMs, and the unmodified com-
pressive strength of the mortar [79]. The use of other materials, such as PVC or other pol-
ymers, should allow the use of inorganic PCMs not compatible with metals, such as salt
hydrates. However, there are almost no studies on them, probably due to several critical
issues such as phase separation, supercooling, and the lack of long-term stability
[58,59,79,80].
Table 7 lists other experimental studies focused on the inclusion of PCMs in concrete
and mortars between 2019 and 2021. With respect to the studies presented in Tables 5 and
6, no salt hydrates are considered, but a greater variety among organic PCMs can be ap-
preciated. In fact, a consistent number of them belongs to alcohols or fatty acids, as well
as paraffins. The majority of PCMs were included in mortars in a shape-stabilized form,
mainly obtained through the impregnation of natural porous materials (e.g., perlite, pum-
ice, diatomite, vermiculite). Other studies evaluated the inclusion of PCMs in microencap-
sulated form, and only a few works investigated the direct inclusion in the mortar [81–
83]. Portland cement is the most studied binder, but alkali-activated cements and geopol-
ymers also seem to be considered as viable alternatives. No problems due to chemical or
physical interactions between PCM and mortar/concrete emerged; nevertheless, the per-
centage of PCM must be limited (up to 20 wt% in the mortars), mainly to control the de-
crease in mechanical properties, such as compressive strength and flexural strength
[82,84–87]. Instead, the decrease in the workability of fresh mortar/concrete is usually
solved by adding more water in the initial mixing of the materials or by using a super-
plasticizer [81,84,88,89].
Polymers 2022, 14, 620 19 of 33
Table 7. Experimental studies on the inclusion of PCMs in concrete and mortars between 2019 and
2021.
PCM Latent
Incorporation in
PCM PCM Tmelt Supporting Building Heat
Reference PCM the Results
Form (°C) Material Material Capacity
Concrete/Mortar
(kJ/kg)
The PCM prevented a concrete
temperature rise, improved its
[81] Butyl stearate bulk 23.4 none concrete direct 134.2 workability, and reduced the
corrosive damages on steel
embedded in concrete.
The PCM lowered both the phase
change temperatures (from 37–40 °C
to 13–17 °C) and the phase change
Hydraulic
enthalpy (from 129 kJ/kg to 7–9
lime, gypsum, impregnated
[84] PEG 1000 SSPCM 37–40 Lecce stone 129 kJ/kg). Both flexural and
cement aggregates
compressive strength showed a
mortars
considerable decrease for all the
binders with a water increment of
15%.
Microencapsulated PCM (up to 20%)
Alkali- enhanced the heat storage capacity
Micronal 5008 acrylic
[85] micro 23 activated microcapsules 100 of the cement but decreased its
(Octadecane) polymer
cements compressive strength from 43% to
50%.
The addition of composite PCM to
the cement led to a significant
Capric acid and decrease in its compressive strength
myristyl alcohol expanded impregnated (from 54% to 82% decrease for 10
[90] SSPCM 17–32 cement 167.2
(weight ratio = perlite aggregates wt% to 30 wt% of composite
9:1) addition). Indoor temperature
fluctuations were reduced, and no
leakage was shown.
The use of steel balls (30%)
prevented leakage problems but
reduced the concrete’s compressive
strength by 18%. A total of 5% of the
blending of cement mass was replaced with slag
[88] Butyl stearate macro 19 steel balls concrete macrocapsules 107.3 and fly ash to solve this problem.
and concrete Coarse aggregates substitution (10%
in volume) with steel balls greatly
enhanced the heat transfer efficiency
without lowering too much the
mechanical properties.
PCM addition to concrete reduced
compressive strength from 18.5 MPa
(0% PCM) to 14.9 MPa (20% PCM).
The addition of PCM up to 10% does
not cause significant changes in
[82] paraffin bulk 20–23 none concrete direct 107.3
flexural and compressive strength.
The direct incorporation of PCM
increased the liquid/binder ratio and
decreased the water absorption of
concrete.
The addition of PCM from 0% to
Alkali- blending of 30% caused a serious reduction in
[91] BSF26 micro 26 not specified activated microcapsules 110 mechanical properties (compressive
cements and cement and flexural strength). No more than
20% of PCM should be added.
The mortar was enriched with 6 wt%
Portland of PCM, enhanced with copper oxide
[83] 1-dodecanol bulk 22 none direct 195
cement and titania, generating a 10%
decrease in the compressive strength
Polymers 2022, 14, 620 20 of 33
In this period (2016–2021), particular attention was put on applications related to spe-
cific seasonal temperatures and variations, leading to a very peculiar situation that re-
quired the use of multiple PCMs (with different phase change temperatures), or well-tai-
lored PCMs, as previously detailed for microencapsulated PCMs. In addition, following
both the climate change issues all over the world and the acceleration of the building sec-
tor in countries (e.g., India and Saudi Arabia) with extremely hot climates (up to 45 °C),
PCM investigation specifically tailored for particular climates is a practical and actual
need [8,19,57,99,100].
In this context, a direct contribution is also given to the environmental aspects,
through a more focused investigation concerning the life cycle assessment analysis per-
formed on PCM applications in constructions. Nevertheless, the environmental impact of
PCMs’ use in the building sector is still an aspect that needs to be studied and that could
have important outcomes on the choice of materials and technologies [101]. For instance,
one of the advantages of macro-encapsulation over micro-encapsulation is the easier sep-
aration between the PCM and the shell material, which simplifies their recovery at the end
of life [56]. This perspective was investigated by Kylili and Fokaides [102], which studied
the impact on the environment through a literature review on the life cycle assessment
(LCA) studies performed on PCM applications in buildings. For the most part, these stud-
ies used the EI99 Environmental Impact Assessment methodology and focused only on
some of the life cycle stages of the investigated materials, particularly the manufacturing,
operational, and disposal stages. Compared to organics (such as paraffins, fatty acids, pol-
yethylene glycol), inorganic PCMs (such as salts hydrates) are less dangerous as they are
chemically stable and can be recycled (Table 8). Other variables were the seasonal condi-
tions, as some studies considered only the summer or the winter season, and the systems’
lifetime, which varied between 50 and 100 years. The results demonstrated that PCMs
incorporated in building materials have a lower impact on the environment compared to
reference cases, without any addition of PCMs. However, these LCA findings are highly
dependent on the scopes and considered phases. The majority of studies did not consider
the lifetime duration of PCM-enriched building materials. Between the manufacture, op-
erational, and disposal stages of these materials, usually only one or two are analyzed.
The results showed that the use of PCMs in building materials is not always environmen-
tally friendly and is highly dependent on life stages. For instance, one of the studies high-
lighted that the addition of PCMs in a ventilated double-skin façade is environmentally
efficient only if the operational period of the façade lasts at least 31 years. Below this value,
the environmental impacts due to manufacturing and the disposal of the façade would
overcome the benefits provided by the addition of PCM [102].
Table 8. Environmental impact of PCMs for building applications. Adaptation from Kylili and
Fokaides [18,101,102].
studies, such as those listed in Table 9, try to use natural materials (e.g., wood) or wastes
(e.g., fly ashes, glass scraps) as supporting materials to produce SSPCMs.
Table 9. Experimental studies on the use of natural substances or wastes as PCMs or supporting
materials.
Composite Latent
PCM Tmelt Supporting Type of
Reference PCM Heat Capacity Remarks and Results
(°C) Material Incorporation
(kJ/kg)
capric acid, vacuum
[105] 25.5 fly ash 45.38 Fly ash was obtained from a power plant.
lauric acid adsorption
transparent The composite material showed good transmittance up to
(delignified) 84% by decreasing thickness (up to 0.5 mm) of the
vacuum
[106] PEG 1000 38 wood and 76 composite material. No changes in elastic modulus were
impregnation
polymethyl observed, except a reduction in flexural strength (70.5 MPa
methacrylate instead of 129.6 MPa) due to the inclusion of PCM.
Heptadecane 62 wt% was found to be the optimum
n- activated carbon one-step
[107] 25.1 138.2 content, to avoid leakage and enhance thermal
heptadecane from pine cones impregnation
conductivity.
mechanical A paraffin/rice husk ratio equal to 50% prevented leakage
[108] paraffin 52.1 rice husk ash mixing and 68.1 problems. PCM and rice husk ash showed good
impregnation compatibility and thermal stability.
The composite showed good chemical and thermal
capric acid
performance stability after 600 phase change cycles. The
(83 wt%)
presence of PCMs decreased the water absorption from
and stearic
Scots pine vacuum 80% to 20%, enhancing wood’s hydrophobicity and anti-
[109] acid (17 24.7 94
sapwood impregnation swelling efficiency. The mechanical properties of wood
wt%)
were also enhanced: modulus of rupture (+22.3%),
eutectic
modulus of elasticity (25.3%), and compression strength
mixture
parallel to grain (24.5%).
NaHPO4 ∗
12H2O (58
wt%) and The composite material, with 40% of diatomite, was coated
diatomite, impregnation,
Na2CO3 ∗ with a polymer to avoid leakage problems. Supercooling
[110] 25 polyurethane coating, UV 102.6
10H2O (42 was almost eliminated (0.5 °C) performance stability
acrylate curing
wt%) confirmed up to 300 phase change cycles.
eutectic
mixture
A microencapsulated PCM was integrated (up to 20 wt%)
Nextek 24D
in a soil and reed fiber mixture. The thermal conductivity
(paraffin
silty-clay soil and mechanical decreased by up to 14%. Water vapor permeability’s
[111] and 22.4 Not specified
reed fiber mixing decrease was 20%. The compressive strength was not
polymeric
affected by the addition of PCM; however, the soil-fiber
shell)
mixture itself showed low values of compressive strength.
PCM and supporting materials are both biodegradable and
obtained from renewable sources. Good chemical
PureTemp cuttlebone one-step 145 for both
[112] 23 compatibility and limited leakage was demonstrated with a
23 pomelo peel impregnation composites
thermal storage efficiency equal to 70% of the pure PCM.
The performance stability was confirmed up to 100 cycles.
The PCM impregnation efficiency in the glass-ceramic
porcelain
vacuum foam was between 24% and 39%. Thermal properties still
[113] organic 29.9 stoneware and Not specified
impregnation need to be measured and leakage problems need to be
soda-lime glass
addressed.
expanded The addition of expanded graphite up to 7% led to a
ultrasonication
OM37 graphite and decrease in latent heat storage capacity, while thermal
[114] 39.1 and vacuum 99.3
(inorganic) expanded conductivity increased by 114% and no leakage was
impregnation
vermiculite detected.
expanded glass
PureTemp vacuum Glass aggregates absorbed up to 80% of PCM and, when
[115] 201 aggregates and 92.7
23 impregnation coated with fly ash, showed no leakage problems.
fly ash (coating)
Polymers 2022, 14, 620 23 of 33
In addition, to make PCM-enhanced building materials more attractive for the mar-
ket, in this last period the need arose to perform comprehensive cost analyses and eco-
nomic evaluations on the applications of these materials in the building sector, as high-
lighted by different authors [8,19,57,79]. Furthermore, to achieve this scope, the adoption
of common and standardized procedures should be considered. Firstly, they should con-
cern all the procedures used to design and produce PCMs-enriched materials for the
building sector, such as mortars. Secondly, the methods used to measure the thermal
properties of PCM-based materials should also be standardized. For example, one of the
most common methods currently used to evaluate the melting temperature and phase
change latent heat is differential scanning calorimetry (DSC), which analyses samples
with a mass in the order of a few milligrams. This amount is not enough to give a good
measure of heterogeneous materials, such as a mixture of cement mortar and micro-en-
capsulated PCM leading to a mismatch between measured and real values of the investi-
gated thermal property [57,79,80]. To enhance the measurements’ standardization, other
methods could be used, such as the T-history method or the use of a heat flow meter ap-
paratus, which is suggested both by the Quality Assurance RAL-GZ 896 [116] and the
Active Standard ASTM C1784 [117] documents.
advantages, the research interest in the development of advanced HGHEs for building
applications has grown in recent years.
The GCHP technology is not new, as its first applications are reported in the first
decade of the twentieth century; nevertheless, the problems exposed previously concern-
ing installation and functioning strongly hindered its wide spreading. For this reason, the
integration of PCMs in vertical and horizontal GHEs is still a recent field of research and
there are few studies concerning this technology, most of which are numerical studies.
Table 10 shows the different studies that analyzed PCMs’ integration in different parts of
a GCHP. Between 2012 and 2015, there were only works dedicated to a numerical analysis
of PCM integration in the GHE backfilling material, while from 2016 to now only a few
studies performed empirical investigations. Several studies aimed to assess which were
the advantages and issues related to PCMs’ addition in the backfilling material.
Dehdezi et al. [124] analyzed the thermal properties of a microencapsulated paraffin
wax-soil mixture and simulated numerically the effects of the PCM addition on the soil
temperature around a shallow-ground GCHP. The addition of microencapsulated paraf-
fin wax caused a reduction in soil thermal conductivity and diffusivity, with a 3 °C lower
temperature variation in the soil and a higher (17%) coefficient of performance (COP) of
the GCHP. Similar findings were highlighted by Bottarelli et al. [125], who studied the
coupling of a novel shallow GHE, a so-called flat-panel, with microencapsulated paraffin
mixed directly with the backfill soil. The presence of microencapsulated paraffin
smoothed the ground temperature variations due to seasonality and GCHP action, also
ensuring a higher COP for the GCHP itself. To mitigate the problem concerning low ther-
mal conductivity, Wang et al. [126] compared, through a numerical simulation, the sepa-
rated use of soil, a mixture of n-decanoic acid and lauric acid (7:3 v:v) as pure PCM, and
PCM with the addition of metal particles, as grout in a vertical GCHP. The results demon-
strated that the replacement of common soil with the n-decanoic acid/lauric acid mixture
could reduce the amount of land area needed to install the GCHP by 18%, while the use
of enhanced PCM led to a reduction of land area of 29%. The importance of thermal con-
ductivity enhancement for the PCM grout was also confirmed by Chen et al. [127], who
investigated paraffin as a PCM. However, thermal conductivity also depends on soil char-
acteristics, such as groundwater presence, which increases it. Qi et al. [128] investigated
the influence of different PCMs (RT27, a commercial paraffin from Rubytherm and a mix-
ture of capric acid and lauric acid (66:34)), with and without the addition of metal parti-
cles, on the backfilling grout of a vertical GCHP, finding that the thermal radius effect was
smaller using PCMs independently of the metal particles’ addition. Zhang et al. [129] com-
pared the performance of a common vertical GCHP with a shallow-ground GCHP sur-
rounded by a cylindrical underground thermal battery (UTB) filled with water and a com-
mercial mixture of salt hydrates as the PCM. The results demonstrated that the UTB guar-
antees more stable temperatures of heat transfer fluid with a ten times shorter borehole
compared to a conventional vertical GCHP. Bonamente and Aquino [130] carried out a
performance analysis, through numerical simulation, and a life cycle assessment study of
a vertical GCHP coupled with an upstream LHTES for space conditioning. The analysis
was based on a prototype and considered the impacts of the system from the production
of input materials to the end-of-life of the GCHP. The LCA analysis highlighted the high
relative impact of the LHTES but also the capacity to lower the whole plant energy con-
sumption and environmental impact.
Since 2016, only a small number of studies have experimentally verified the effect of
PCMs in the backfilling material of a GCHP. Li et al. [131] simulated the use of crushed
stone and shape-stabilized PCM (a mixture of decanoic acid and lauric acid (3:2 v:v)) con-
crete enhanced with silica and graphite as grout for a U-tube vertical GHE. The model’s
effectiveness was proven through an experimental measurement of the backfilling mate-
rial at a laboratory scale. The simulation proved that PCMs increase the heat storage ca-
pacity of the borehole and lower its influence radius on the temperature of surrounding
soil, while graphite and silica enhance the heat exchange. Yang et al. [132] performed an
Polymers 2022, 14, 620 25 of 33
Table 10. Studies on the integration of PCMs in different parts of GCHPs. Num = numerical study,
exp = experimental study, H = horizontal GHE, V = vertical GHE.
Melting Latent Heat
Reference Study Type GHE PCM PCM Employment
Point (°C) (kJ/kg)
one-dimensional finite difference micro-encapsulated
[124] num H 26 160 backfilling soil
transient heat transport model paraffin
decanoic acid and GHE’s borehole as
[126] num three-dimensional V 25 /
lauric acid mixture. grout.
water, micro-
[125] num two-dimensional H 0, 26 / backfilling soil
encapsulated paraffin
hydrate sodium sulfate
[134] num modified composite model V 8.3 95.4 GCHP
(type 47)
computational fluid dynamics RT6 8.0–8.5 140
[137] num V GCHP
simulations RT27 25.0–25.5 146
three-dimensional unsteady decanoic acid and backfill material in
[131] num V 20.15 128
model lauric acid mixture a GCHP
paraffin RT27.
three-dimensional finite element decanoic acid and 28–30 179 GHE’s borehole
[128] num V
model lauric acid mixture 20.4 138.8 grout.
(66:34)
micro-encapsulated
[136] exp / V 39.5 9.0–20.9 HTF in a GCHP
methyl stearate
[138] num finite element model V paraffin / 190 GCHP borehole
micro-encapsulated
paraffin.
three-dimensional unsteady 23–27 150
[139] num V Shape-stabilized grout with a GCHP
model 19.9 109.2
decanoic and lauric
acid mixture
Polymers 2022, 14, 620 26 of 33
From these results, it is clear that very different ways of enhancing GCHPs’ perfor-
mances and issues through PCMs have been tested, but current studies suffer from a lack
of experimental data and validation, as almost all of them present only numerical anal-
yses. Another important aspect is the system’s end-of-life, as PCM recovery and disposal
would be easier for horizontal GCHPs compared to vertical GCHPs.
Another possible use of PCMs concerns energy piles technology, where a closed loop
consisting of a GHE integrated into the concrete of foundation piles of structures can be
exploited [140]. Although several studies have already analyzed this technology, energy
piles still face different critical challenges, such as the effect of temperature variations, due
to the GCHP, on the internal stress distribution of the concrete pile or its bearing capacity
[141].
In addition, the integration of PCMs into concrete would enhance the GCHP’s effi-
ciency. However, there are many challenges linked to interactions between concrete, GHE
pipes, and PCMs, while the number of related studies is still very limited [142,143]. The
integration between PCMs and backfilling materials, or concrete in energy piles, could
also take advantage of previous studies performed in the building sector. For instance,
studies on the integration of PCMs in envelope materials as mortars or concrete (Table 7)
may be a good starting point to create backfilling grouts with known thermal and me-
chanical properties. However, GHE backfilling contributes to the building’s footprint.
Therefore, its enhancement should be considered as a further strategic opportunity for the
integration of larger thermal energy storage, which can better mitigate the synchrony loss
between supply and demand in using renewable energies.
Furthermore, the integration of PCMs with GHE backfilling materials does not affect
the building volume, (as occurs with HVAC), and the lower digging cost compensates for
its installation. Recently, to further reduce the GCHPs’ installation costs by shortening the
GHEs’ length, multi-source heat pump systems have been studied and designed on a res-
idential scale and for small building volumes [144,145]. The focus is on the exploitation of
the most advantageous energy source (air, ground, solar), which is selected according to
evolved temperatures. However, the ground is the sole source able to carry out thermal
storage and can therefore play a pivot role, which may be improved by PCMs.
4. Conclusions
This work analyzed the main review and experimental studies concerning the inte-
gration of phase change materials into building applications during the last twenty years.
The evolution of the research in this field was divided into three main stages (2000–2010,
2011–2015, and 2016–2021), and the evolution concerning phase change materials and
their application was discussed, with a focus on the studies performed between 2019 and
2021. Some conclusions can be drawn:
- among different classes of PCM, paraffins are generally preferred for integration in
building materials, due to their high performance/cost ratio, but more recent studies
(2019–2021) also consider other organic PCMs, namely alcohols and fatty acids, as
more specific applications need tailored materials.
Polymers 2022, 14, 620 27 of 33
Author Contributions: Conceptualization, M.M. and M.B.; investigation, S.B., F.B. and S.M.; writ-
ing—original draft preparation, S.B., F.B. and S.M. (Simona Marinelli).; writing—review and edit-
ing, S.B., F.B. and S.M. (Sebastiano Merchiori); supervision, B.R. and M.M.; project administration,
M.B.; funding acquisition, M.B. All authors have read and agreed to the published version of the
manuscript.
Funding: This work was supported by the Emilia Romagna region in the framework of the project
“CLIWAX” co-financed by 2014–2020 POR FESR Emilia-Romagna Region Italy DGR 774/2015—
CUP F71F18000160009.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. United Nations Environment Programme (UNEP); International Energy Agency (IEA). Towards a Zero-Emission, Efficient, and
Resilient Buildings and Construction Sector; International Energy Agency: Paris, France, 2017.
2. International Energy Agency (IEA). Tracking Buildings 2020; International Energy Agency: Paris, France, 2020; pp. 1–11.
3. European Parliament. Directive (EU) 2018/844 of the European Parliament and of the Council of 30 May 2018 amending. Off. J.
Eur. Union 2018, 156, 75–91.
4. European Parliament. Directive 2010/31/EU of the European Parliament and of the Council of 19 May 2010 on the Energy Per-
formance of Buildings. Off. J. Eur. Union 2010, 153, 23.
5. Stritih, U.; Tyagi, V.V.; Stropnik, R.; Paksoy, H.; Haghighat, F.; Joybari, M.M. Integration of passive PCM technologies for net-
zero energy buildings. Sustain. Cities Soc. 2018, 41, 286–295. https://doi.org/10.1016/j.scs.2018.04.036.
6. Sharma, A.; Tyagi, V.V.; Chen, C.R.; Buddhi, D. Review on thermal energy storage with phase change materials and applica-
tions, Renew. Sustain. Energy Rev. 2009, 13, 318–345. https://doi.org/10.1016/j.rser.2007.10.005.
Polymers 2022, 14, 620 28 of 33
7. Jemmal, Y.; Zari, N.; Maaroufi, M. Thermophysical and chemical analysis of gneiss rock as low cost candidate material for
thermal energy storage in concentrated solar power plants. Sol. Energy Mater. Sol. Cells 2016, 157, 377–382.
https://doi.org/10.1016/j.solmat.2016.06.002.
8. Rathore, P.K.S.; Shukla, S.K. Potential of macroencapsulated PCM for thermal energy storage in buildings: A comprehensive
review. Constr. Build. Mater. 2019, 225, 723–744. https://doi.org/10.1016/j.conbuildmat.2019.07.221.
9. Baetens, R.; Jelle, B.P.; Gustavsen, A. Phase change materials for building applications: A state-of-the-art review. Energy Build.
2010, 42, 1361–1368. https://doi.org/10.1016/j.enbuild.2010.03.026.
10. Osterman, E.; Tyagi, V.V.; Butala, V.; Rahim, N.A.; Stritih, U. Review of PCM based cooling technologies for buildings. Energy
Build. 2012, 49, 37–49. https://doi.org/10.1016/j.enbuild.2012.03.022.
11. Kenisarin, M.; Mahkamov, K. Salt hydrates as latent heat storage materials: Thermophysical properties and costs. Sol. Energy
Mater. Sol. Cells 2015, 145, 1–32. https://doi.org/10.1016/j.solmat.2015.10.029.
12. Wong-Pinto, L.-S.; Milian, Y.; Ushak, S. Progress on use of nanoparticles in salt hydrates as phase change materials. Renew.
Sustain. Energy Rev. 2020, 122, 109727. https://doi.org/10.1016/j.rser.2020.109727.
13. Lin, Y.; Alva, G.; Fang, G. Review on thermal performances and applications of thermal energy storage systems with inorganic
phase change materials. Energy 2018, 165, 685–708. https://doi.org/10.1016/j.energy.2018.09.128.
14. Farid, M.M.; Khudhair, A.M.; Razack, S.A.K.; Al-Hallaj, S. A review on phase change energy storage: Materials and applications.
Energy Convers. Manag. 2004, 45, 1597–1615. https://doi.org/10.1016/j.enconman.2003.09.015.
15. Mohamed, S.A.; Al-Sulaiman, F.A.; Ibrahim, N.I.; Zahir, H.M.; Al-Ahmed, A.; Saidur, R.; Yılbaş, B.S.; Sahin, A.Z. A review on
current status and challenges of inorganic phase change materials for thermal energy storage systems. Renew. Sustain. Energy
Rev. 2017, 70, 1072–1089. https://doi.org/10.1016/j.rser.2016.12.012.
16. Nazir, H.; Batool, M.; Bolivar Osorio, F.J.; Isaza-Ruiz, M.; Xu, X.; Vignarooban, K.; Phelan, P.; Inamuddin; Kannan, A.M. Recent
developments in phase change materials for energy storage applications: A review. Int. J. Heat Mass Transf. 2019, 129, 491–523.
https://doi.org/10.1016/j.ijheatmasstransfer.2018.09.126.
17. Fallahi, A.; Guldentops, G.; Tao, M.; Granados-Focil, S.; Van Dessel, S. Review on solid-solid phase change materials for thermal
energy storage: Molecular structure and thermal properties. Appl. Therm. Eng. 2017, 127, 1427–1441. https://doi.org/10.1016/j.ap-
plthermaleng.2017.08.161.
18. de Gracia, A.; Cabeza, L.F. Phase change materials and thermal energy storage for buildings. Energy Build. 2015, 103, 414–419.
https://doi.org/10.1016/j.enbuild.2015.06.007.
19. Akeiber, H.; Nejat, P.; Majid, M.Z.A.; Wahid, M.A.; Jomehzadeh, F.; Zeynali Famileh, I.; Calautit, J.K.; Hughes, B.R.; Zaki, S.A.
A review on phase change material (PCM) for sustainable passive cooling in building envelopes. Renew. Sustain. Energy Rev.
2016, 60, 1470–1497. https://doi.org/10.1016/j.rser.2016.03.036.
20. Iten, M.; Liu, S.; Shukla, A. A review on the air-PCM-TES application for free cooling and heating in the buildings. Renew.
Sustain. Energy Rev. 2016, 61, 175–186. https://doi.org/10.1016/j.rser.2016.03.007.
21. Aranda-Usón, A.; Ferreira, G.; López-Sabirón, A.M.; Mainar-Toledo, M.D.; Zabalza Bribián, I. Phase change material applica-
tions in buildings: An environmental assessment for some Spanish climate severities. Sci. Total Environ. 2013, 444, 16–25.
https://doi.org/10.1016/j.scitotenv.2012.11.012.
22. de Gracia, A.; Rincón, L.; Castell, A.; Jimenez, M.; Boer, D.; Medrano, M.; Cabeza, L.F. Life Cycle Assessment of the inclusion of
phase change materials (PCM) in experimental buildings. Energy Build. 2010, 42, 1517–1523.
https://doi.org/10.1016/j.enbuild.2010.03.022.
23. Baldassarri, C.; Sala, S.; Caverzan, A.; Lomperti Tornaghi, M. Environmental and spatial assessment for the ecodesign of a
cladding system with embedded Phase Change Materials. Energy Build. 2017, 156, 374–389.
https://doi.org/10.1016/j.enbuild.2017.09.011.
24. Nienborg, B.; Gschwander, S.; Munz, G.; Fröhlich, D.; Helling, T.; Horn, R.; Weinläder, H.; Klinker, F.; Schossig, P. Life Cycle
Assessment of thermal energy storage materials and components. Energy Procedia 2018, 155, 111–120.
https://doi.org/10.1016/j.egypro.2018.11.063.
25. Sutterlin, W.R. Phase Change Materials: A Brief Comparison of Ice Packs, Salts, Paraffins, and Vegetable-Derived Phase Change
Materials—Pharmaceutical Outsourcing. Available online: https://www.pharmoutsourcing.com/Featured-Articles/37854-
Phase-Change-Materials-A-Brief-Comparison-of-Ice-Packs-Salts-Paraffins-and-Vegetable-derived-Phase-Change-Materials/
(accessed on 2022.02.04).
26. Zalba, B.; Marı́n, 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. https://doi.org/10.1016/S1359-4311(02)00192-8.
27. Verma, P.; Varun; Singal, S. Review of mathematical modeling on latent heat thermal energy storage systems using phase-
change material. Renew. Sustain. Energy Rev. 2008, 12, 999–1031. https://doi.org/10.1016/j.rser.2006.11.002.
28. 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. https://doi.org/10.1016/j.rser.2007.06.009.
29. Jegadheeswaran, S.; Pohekar, S.D. Performance enhancement in latent heat thermal storage system: A review. Renew. Sustain.
Energy Rev. 2009, 13, 2225–2244. https://doi.org/10.1016/j.rser.2009.06.024.
30. Tyagi, V.V.; Buddhi, D. PCM thermal storage in buildings: A state of art. Renew. Sustain. Energy Rev. 2007, 11, 1146–1166.
https://doi.org/10.1016/j.rser.2005.10.002.
Polymers 2022, 14, 620 29 of 33
31. Chekhovskoi, V.Y. Thermal expansion and density of 80.5% LiF-19.5% CaF2 eutectic. High Temp. 2000, 38, 197–202.
https://doi.org/10.1007/BF02755945.
32. Agyenim, F.; Hewitt, N.; Eames, P.; Smyth, M. A review of materials, heat transfer and phase change problem formulation for
latent heat thermal energy storage systems (LHTESS). Renew. Sustain. Energy Rev. 2010, 14, 615–628.
https://doi.org/10.1016/j.rser.2009.10.015.
33. Raj, V.A.A.; Velraj, R. Review on free cooling of buildings using phase change materials. Renew. Sustain. Energy Rev. 2010, 14,
2819–2829. https://doi.org/10.1016/j.rser.2010.07.004.
34. Zalba, B.; Marı́n, J.M.; Cabeza, L.F.; Mehling, H. Free-cooling of buildings with phase change materials. Int. J. Refrig. 2004, 27,
839–849. https://doi.org/10.1016/j.ijrefrig.2004.03.015.
35. Zalba, B.C.L.; Marin, J.M.; Sanchez-Valverde, B. Free cooling: An application of PCMS in TES. In Proceedings of the 3rd Work-
shop IEA ECES IA (Annex 17), Tokyo, Japan, 1–2 October 2002.
36. Lazaro, A.; Dolado, P.; Marín, J.M.; Zalba, B. PCM–air heat exchangers for free-cooling applications in buildings: Experimental
results of two real-scale prototypes. Energy Convers. Manag. 2009, 50, 439–443. https://doi.org/10.1016/j.enconman.2008.11.002.
37. Agyenim, F.; Eames, P.; Smyth, M. Heat transfer enhancement in medium temperature thermal energy storage system using a
multitube heat transfer array. Renew. Energy 2010, 35, 198–207. https://doi.org/10.1016/j.renene.2009.03.010.
38. Papanicolaou, E.; Belessiotis, V. Transient natural convection in a cylindrical enclosure at high Rayleigh numbers. Int. J. Heat
Mass Transf. 2002, 45, 1425–1444. https://doi.org/10.1016/S0017-9310(01)00258-7.
39. Zhang, P.; Ma, Z.; Wang, R. An overview of phase change material slurries: MPCS and CHS. Renew. Sustain. Energy Rev. 2010,
14, 598–614. https://doi.org/10.1016/j.rser.2009.08.015.
40. Lacroix, M. Study of the heat transfer behavior of a latent heat thermal energy storage unit with a finned tube. Int. J. Heat Mass
Transf. 1993, 36, 2083–2092. https://doi.org/10.1016/S0017-9310(05)80139-5.
41. Jegadheeswaran, S.; Pohekar, S.; Kousksou, T. Exergy based performance evaluation of latent heat thermal storage system: A
review. Renew. Sustain. Energy Rev. 2010, 14, 2580–2595. https://doi.org/10.1016/j.rser.2010.07.051.
42. Lock, G.S. Latent Heat Transfer; Oxford University Press: Oxford, UK, 1994.
43. Cabeza, L.F.; Castell, A.; Barreneche, C.; De Gracia, A.; Fernández, A.I. Materials used as PCM in thermal energy storage in
buildings: A review. Renew. Sustain. Energy Rev. 2011, 15, 1675–1695. https://doi.org/10.1016/j.rser.2010.11.018.
44. Zhou, D.; Zhao, C.Y.; Tian, Y. Review on thermal energy storage with phase change materials (PCMs) in building applications.
Appl. Energy 2012, 92, 593–605. https://doi.org/10.1016/j.apenergy.2011.08.025.
45. Kalnæs, S.E.; Jelle, B.P. Phase change materials and products for building applications: A state-of-the-art review and future
research opportunities. Energy Build. 2015, 94, 150–176. https://doi.org/10.1016/j.enbuild.2015.02.023.
46. 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. https://doi.org/10.1016/j.rser.2012.10.022.
47. Konuklu, Y.; Ostry, M.; Paksoy, H.O.; Charvat, P. Review on using microencapsulated phase change materials (PCM) in build-
ing applications. Energy Build. 2015, 106, 134–155. https://doi.org/10.1016/j.enbuild.2015.07.019.
48. Cao, L.; Su, D.; Tang, Y.; Fang, G.; Tang, F. Properties evaluation and applications of thermal energystorage materials in build-
ings. Renew. Sustain. Energy Rev. 2015, 48, 500–522. https://doi.org/10.1016/j.rser.2015.04.041.
49. Pomianowski, M.; Heiselberg, P.; Zhang, Y. Review of thermal energy storage technologies based on PCM application in build-
ings. Energy Build. 2013, 67, 56–69. https://doi.org/10.1016/j.enbuild.2013.08.006.
50. Soares, N.; Costa, J.J.; Gaspar, A.R.; Santos, P. Review of passive PCM latent heat thermal energy storage systems towards
buildings’ energy efficiency. Energy Build. 2013, 59, 82–103. https://doi.org/10.1016/j.enbuild.2012.12.042.
51. Memon, S.A. Phase change materials integrated in building walls: A state of the art review. Renew. Sustain. Energy Rev. 2014, 31,
870–906. https://doi.org/10.1016/j.rser.2013.12.042.
52. Kuznik, F.; David, D.; Johannes, K.; Roux, J.-J. A review on phase change materials integrated in building walls. Renew. Sustain.
Energy Rev. 2011, 15, 379–391. https://doi.org/10.1016/j.rser.2010.08.019.
53. Ling, T.C.; Poon, C.S. Use of phase change materials for thermal energy storage in concrete: An overview. Constr. Build. Mater.
2013, 46, 55–62. https://doi.org/10.1016/j.conbuildmat.2013.04.031.
54. Castell, A.; Menoufi, K.; de Gracia, A.; Rincón, L.; Boer, D.; Cabeza, L.F. Life Cycle Assessment of alveolar brick construction
system incorporating phase change materials (PCMs). Appl. Energy 2013, 101, 600–608. https://doi.org/10.1016/j.apen-
ergy.2012.06.066.
55. Menoufi, K.; Castell, A.; Farid, M.; Boer, D.; Cabeza, L.F. Life Cycle Assessment of experimental cubicles including PCM man-
ufactured from natural resources (esters): A theoretical study. Renew. Energy 2013, 51, 398–403.
https://doi.org/10.1016/j.renene.2012.10.010.
56. Liu, Z.; Yu, Z.; Yang, T.; Qin, D.; Li, S.; Zhang, G.; Haghighat, F.; Joybari, M.M. A review on macro-encapsulated phase change
material for building envelope applications. Build. Environ. 2018, 144, 281–294. https://doi.org/10.1016/j.buildenv.2018.08.030.
57. Singh Rathore, P.K.; Shukla, S.K.; Gupta, N.K. Potential of microencapsulated PCM for energy savings in buildings: A critical
review. Sustain. Cities Soc. 2020, 53, 101884. https://doi.org/10.1016/j.scs.2019.101884.
58. da Cunha, S.R.L.; de Aguiar, J.L.B. Phase change materials and energy efficiency of buildings: A review of knowledge. J. Energy
Storage 2020, 27, 101083. https://doi.org/10.1016/j.est.2019.101083.
59. Song, M.; Niu, F.; Mao, N.; Hu, Y.; Deng, S.S. Review on building energy performance improvement using phase change mate-
rials. Energy Build. 2018, 158, 776–793. https://doi.org/10.1016/j.enbuild.2017.10.066.
Polymers 2022, 14, 620 30 of 33
60. Zhu, N.; Li, S.; Hu, P.; Wei, S.; Deng, R.; Lei, F. A review on applications of shape-stabilized phase change materials embedded
in building enclosure in recent ten years. Sustain. Cities Soc. 2018, 43, 251–264. https://doi.org/10.1016/j.scs.2018.08.028.
61. He, W.; Yu, C.; Yang, J.; Yu, B.; Hu, Z.; Shen, D.; Liu, X.; Qin, M.; Chen, H. Experimental study on the performance of a novel
RC-PCM-wall. Energy Build. 2019, 199, 297–310. https://doi.org/10.1016/j.enbuild.2019.07.001.
62. Li, Y.; Darkwa, J.; Kokogiannakis, G.; Su, W. Phase change material blind system for double skin façade integration: System
development and thermal performance evaluation. Appl. Energy 2019, 252, 113376. https://doi.org/10.1016/j.apen-
ergy.2019.113376.
63. Meng, E.; Wang, J.; Yu, H.; Cai, R.; Chen, Y.; Zhou, B. Experimental study of the thermal protection performance of the high
reflectivity-phase change material (PCM) roof in summer. Build. Environ. 2019, 164, 106381. https://doi.org/10.1016/j.build-
env.2019.106381.
64. Piselli, C.; Castaldo, V.L.; Pisello, A.L. How to enhance thermal energy storage effect of PCM in roofs with varying solar reflec-
tance: Experimental and numerical assessment of a new roof system for passive cooling in different climate conditions. Sol.
Energy 2019, 192, 106–119. https://doi.org/10.1016/j.solener.2018.06.047.
65. Qiao, Y.; Yang, L.; Bao, J.; Liu, Y.; Liu, J. Reduced-scale experiments on the thermal performance of phase change material
wallboard in different climate conditions. Build. Environ. 2019, 160, 106191. https://doi.org/10.1016/j.buildenv.2019.106191.
66. Valizadeh, S.; Ehsani, M.; Torabi Angji, M. Development and thermal performance of wood-HPDE- PCM nanocapsule floor for
passive cooling in building. Energy Sources Part A Recovery Util. Environ. Eff. 2019, 41, 2114–2127.
https://doi.org/10.1080/15567036.2018.1550125.
67. Alqahtani, T.; Mellouli, S.; Bamasag, A.; Askri, F.; Phelan, P.E. Experimental and numerical assessment of using coconut oil as
a phase-change material for unconditioned buildings. Int. J. Energy Res. 2020, 44, 5177–5196. https://doi.org/10.1002/er.5176.
68. Maleki, B.; Khadang, A.; Maddah, H.; Alizadeh, M.; Kazemian, A.; Ali, H.M. Development and thermal performance of nano-
encapsulated PCM/ plaster wallboard for thermal energy storage in buildings. J. Build. Eng. 2020, 32, 101727.
https://doi.org/10.1016/j.jobe.2020.101727.
69. Musiał, M. Experimental and Numerical Analysis of the Energy Efficiency of Transparent Partitions with a Thermal Storage
Unit. J. Ecol. Eng. 2020, 21, 201–211. https://doi.org/10.12911/22998993/123831.
70. Sonnick, S.; Erlbeck, L.; Gaedtke, M.; Wunder, F.; Mayer, C.; Krause, M.J.; Nirschl, H.; Rädle, M. Passive room conditioning
using phase change materials—Demonstration of a long-term real size experiment. Int. J. Energy Res. 2020, 44, 7047–7056.
https://doi.org/10.1002/er.5406.
71. Yu, H.; Li, C.; Zhang, K.; Tang, Y.; Song, Y.; Wang, M. Preparation and thermophysical performance of diatomite-based com-
posite PCM wallboard for thermal energy storage in buildings. J. Build. Eng. 2020, 32, 101753.
https://doi.org/10.1016/j.jobe.2020.101753.
72. Zhang, Y.; Sun, X.; Medina, M.A. Experimental evaluation of structural insulated panels outfitted with phase change materials.
Appl. Therm. Eng. 2020, 178, 115454. https://doi.org/10.1016/j.applthermaleng.2020.115454.
73. Bolteya, A.M.; Elsayad, M.A.; Belal, A.M. Thermal efficiency of PCM filled double glazing units in Egypt. Ain Shams Eng. J. 2021,
12, 1523–1534. https://doi.org/10.1016/j.asej.2020.12.004.
74. Sinka, M.; Bajare, D.; Jakovics, A.; Ratnieks, J.; Gendelis, S.; Tihana, J. Experimental testing of phase change materials in a warm-
summer humid continental climate. Energy Build. 2019, 195, 205–215. https://doi.org/10.1016/j.enbuild.2019.04.030.
75. Yun, B.Y.; Yang, S.; Cho, H.M.; Wi, S.; Kim, S. Thermal Storage Effect Analysis of Floor Heating Systems Using Latent Heat
Storage Sheets. Int. J. Precis. Eng. Manuf. Green Technol. 2019, 6, 799–807. https://doi.org/10.1007/s40684-019-00131-3.
76. Bogatu, D.I.; Kazanci, O.B.; Olesen, B.W. An experimental study of the active cooling performance of a novel radiant ceiling
panel containing phase change material (PCM). Energy Build. 2021, 243, 110981. https://doi.org/10.1016/j.enbuild.2021.110981.
77. Faraj, K.; Khaled, M.; Faraj, J.; Hachem, F.; Castelain, C. Experimental Study on the Use of Enhanced Coconut Oil and Paraffin
Wax Phase Change Material in Active Heating Using Advanced Modular Prototype. J. Energy Storage 2021, 41, 102815.
https://doi.org/10.1016/j.est.2021.102815.
78. Guo, J.; Dong, J.; Wang, H.; Jiang, Y.; Tao, J. On-site measurement of the thermal performance of a novel ventilated thermal
storage heating floor in a nearly zero energy building. Build. Environ. 2021, 201, 107993. https://doi.org/10.1016/j.build-
env.2021.107993.
79. Rao, V.V.; Parameshwaran, R.; Ram, V.V. PCM-mortar based construction materials for energy efficient buildings: A review on
research trends. Energy Build. 2018, 158, 95–122. https://doi.org/10.1016/j.enbuild.2017.09.098.
80. Berardi, U.; Gallardo, A. Properties of concretes enhanced with phase change materials for building applications. Energy Build.
2019, 199, 402–414. https://doi.org/10.1016/j.enbuild.2019.07.014.
81. Cellat, K.; Tezcan, F.; Kardaş, G.; Paksoy, H. Comprehensive investigation of butyl stearate as a multifunctional smart concrete
additive for energy-efficient buildings. Int. J. Energy Res. 2019, 43, 7146–7158. https://doi.org/10.1002/er.4740.
82. Cunha, S.; Leite, P.; Aguiar, J. Characterization of innovative mortars with direct incorporation of phase change materials. J.
Energy Storage 2020, 30, 101439. https://doi.org/10.1016/j.est.2020.101439.
83. Parameshwaran, R.; Kumar, G.N.; Ram, V.V. Experimental analysis of hybrid nanocomposite-phase change material embedded
cement mortar for thermal energy storage. J. Build. Eng. 2020, 30, 101297. https://doi.org/10.1016/j.jobe.2020.101297.
84. Frigione, M.; Lettieri, M.; Sarcinella, A.; De Aguiar, J.L.B. Applications of Sustainable Polymer-Based Phase Change Materials
in Mortars Composed by Different Binders. Materials 2019, 12, 3502. https://doi.org/10.3390/ma12213502.
Polymers 2022, 14, 620 31 of 33
85. Giro-Paloma, J.; Barreneche, C.; Maldonado-Alameda, A.; Royo, M.; Formosa, J.; Fernandez, A.I.; Chimenos, J.M. Alkali-Acti-
vated Cements for TES Materials in Buildings’ Envelops Formulated With Glass Cullet Recycling Waste and Microencapsulated
Phase Change Materials. Materials 2019, 12, 2144. https://doi.org/10.3390/ma12132144.
86. Hekimoğlu, G.; Nas, M.; Ouikhalfan, M.; Sarı, A.; Kurbetci, Ş.; Tyagi, V.V.; Sharma, R.K.; Saleh, T.A. Thermal management
performance and mechanical properties of a novel cementitious composite containing fly ash/lauric acid-myristic acid as form-
stable phase change material. Constr. Build. Mater. 2020, 274, 122105. https://doi.org/10.1016/j.conbuildmat.2020.122105.
87. Hekimoğlu, G.; Nas, M.; Ouikhalfan, M.; Sarı, A.; Tyagi, V.V.; Sharma, R.K.; Kurbetci, Ş.; Saleh, T.A. Silica fume/capric acid-
stearic acid PCM included-cementitious composite for thermal controlling of buildings: Thermal energy storage and mechanical
properties. Energy 2020, 219, 119588. https://doi.org/10.1016/j.energy.2020.119588.
88. Chang, H.; Jin, L. Preparation and Heat Transfer Performance of Steel Ball Phase Change Concrete. J. New Mater. Electrochem.
Syst. 2020, 23, 204–212. https://doi.org/10.14447/jnmes.v23i3.a08.
89. Shi, J.; Li, M. Lightweight mortar with paraffin/expanded vermiculite-diatomite composite phase change materials: Develop-
ment, characterization and year-round thermoregulation performance. Sol. Energy 2021, 220, 331–342.
https://doi.org/10.1016/j.solener.2021.03.053.
90. Zhang, X.; Xu, M.; Liu, L.; Huan, C.; Zhao, Y.; Qi, C.; Song, K.-I. Experimental study on thermal and mechanical properties of
cemented paste backfill with phase change material. J. Mater. Res. Technol. 2019, 9, 1–12.
https://doi.org/10.1016/j.jmrt.2019.12.047.
91. Kheradmand, M.; Abdollahnejad, Z.; Pacheco-Torgal, F. Alkali-activated cement-based binder mortars containing phase change
materials (PCMs): Mechanical properties and cost analysis. Eur. J. Environ. Civ. Eng. 2020, 24, 1068–1090.
https://doi.org/10.1080/19648189.2018.1446362.
92. Sarı, A.; Tyagi, V.V. Thermal energy storage properties and lab-scale thermal performance in cementitious plaster of composite
phase change material for energy efficiency of buildings. Environ. Prog. Sustain. Energy 2020, 39, e13455.
https://doi.org/10.1002/ep.13455.
93. Sarı, A.; Hekimoğlu, G.; Tyagi, V.; Sharma, R. Evaluation of pumice for development of low-cost and energy-efficient composite
phase change materials and lab-scale thermoregulation performances of its cementitious plasters. Energy 2020, 207, 118242.
https://doi.org/10.1016/j.energy.2020.118242.
94. Valentini, F.; Morandini, F.; Bergamo, M.; Dorigato, A. Development of eco-sustainable plasters with thermal energy storage
capability. J. Appl. Phys. 2020, 128, 1–12. https://doi.org/10.1063/5.0012139.
95. Hasanabadi, S.; Sadrameli, S.M.; Sami, S. Preparation, characterization and thermal properties of surface-modified expanded
perlite/paraffin as a form-stable phase change composite in concrete. J. Therm. Anal. Calorim. 2021, 144, 61–69.
https://doi.org/10.1007/s10973-020-09440-1.
96. Illampas, R.; Rigopoulos, I.; Ioannou, I. Influence of microencapsulated Phase Change Materials (PCMs) on the properties of
polymer modified cementitious repair mortar. J. Build. Eng. 2021, 40, 102328. https://doi.org/10.1016/j.jobe.2021.102328.
97. Ramakrishnan, S.; Pasupathy, K.; Sanjayan, J. Synthesis and properties of thermally enhanced aerated geopolymer concrete
using form-stable phase change composite. J. Build. Eng. 2021, 40, 102756. https://doi.org/10.1016/j.jobe.2021.102756.
98. Shen, Y.; Liu, S.; Zeng, C.; Zhang, Y.; Li, Y.; Han, X.; Yang, L.; Yang, X. Experimental thermal study of a new PCM-concrete
thermal storage block (PCM-CTSB). Constr. Build. Mater. 2021, 293, 123540. https://doi.org/10.1016/j.conbuildmat.2021.123540.
99. Wang, M.; Liu, L.; Zhang, X.-Y.; Chen, L.; Wang, S.-Q.; Jia, Y.-H. Experimental and numerical investigations of heat transfer and
phase change characteristics of cemented paste backfill with PCM. Appl. Therm. Eng. 2019, 150, 121–131.
https://doi.org/10.1016/j.applthermaleng.2018.12.103.
100. Dardir, M.; Panchabikesan, K.; Haghighat, F.; El Mankibi, M.; Yuan, Y. Opportunities and challenges of PCM-to-air heat ex-
changers (PAHXs) for building free cooling applications—A comprehensive review. J. Energy Storage 2019, 22, 157–175.
https://doi.org/10.1016/j.est.2019.02.011.
101. De Falco, M.; Capocelli, M.; Losito, G.; Piemonte, V. LCA perspective to assess the environmental impact of a novel PCM-based
cold storage unit for the civil air conditioning. J. Clean. Prod. 2017, 165, 697–704. https://doi.org/10.1016/j.jclepro.2017.07.153.
102. Kylili, A.; Fokaides, P.A. Life Cycle Assessment (LCA) of Phase Change Materials (PCMs) for building applications: A review.
J. Build. Eng. 2016, 6, 133–143. https://doi.org/10.1016/j.jobe.2016.02.008.
103. Thaib, R.; Hamdani, H.; Amin, M. Utilization of Beeswax/Bentonite as energy storage material on building wall composite. J.
Phys. Conf. Ser. 2019, 1402, 044038. https://doi.org/10.1088/1742-6596/1402/4/044038.
104. Hakim, I.I.; Putra, N.; Agustin, P.D. Measurement of PCM-concrete composites thermal properties for energy conservation in
building material. AIP Conf. Proc. 2020, 2255, 030066. https://doi.org/10.1063/5.0014120.
105. Ji, R.; Zou, Z.; Liu, L.; Wei, S.; Qu, S. Development and energy evaluation of phase change material composite for building
energy-saving. Int. J. Energy Res. 2019, 43, 8674–8683. https://doi.org/10.1002/er.4867.
106. Montanari, C.; Li, Y.; Chen, H.; Yan, M.; Berglund, L.A. Transparent Wood for Thermal Energy Storage and Reversible Optical
Transmittance. ACS Appl. Mater. Interfaces 2019, 11, 20465–20472. https://doi.org/10.1021/acsami.9b05525.
107. Sarabandi, D.; Roudini, G.; Barahuie, F. Activated carbon derived from pine cone as a framework for the preparation of n-
heptadecane nanocomposite for thermal energy storage. J. Energy Storage 2019, 24, 100795.
https://doi.org/10.1016/j.est.2019.100795.
Polymers 2022, 14, 620 32 of 33
108. Liu, Y.; Yu, K.; Lu, S.; Wang, C.; Li, X.; Yang, Y. Experimental research on an environment-friendly form-stable phase change
material incorporating modified rice husk ash for thermal energy storage. J. Energy Storage 2020, 31, 101599.
https://doi.org/10.1016/j.est.2020.101599.
109. Temiz, A.; Hekimoğlu, G.; Köse Demirel, G.; Sarı, A.; Mohamad Amini, M.H. Phase change material impregnated wood for
passive thermal management of timber buildings. Int. J. Energy Res. 2020, 44, 10495–10505. https://doi.org/10.1002/er.5679.
110. Xie, N.; Niu, J.; Zhong, Y.; Gao, X.; Zhang, Z.; Fang, Y. Development of polyurethane acrylate coated salt hydrate/diatomite
form-stable phase change material with enhanced thermal stability for building energy storage. Constr. Build. Mater. 2020, 259,
119714. https://doi.org/10.1016/j.conbuildmat.2020.119714.
111. Alassaad, F.; Touati, K.; Levacher, D.; Sebaibi, N. Impact of phase change materials on lightened earth hygroscopic, thermal
and mechanical properties. J. Build. Eng. 2021, 41, 102417. https://doi.org/10.1016/j.jobe.2021.102417.
112. Biesuz, M.; Valentini, F.; Bortolotti, M.; Zambotti, A.; Cestari, F.; Bruni, A.; Sglavo, V.M.; Sorarù, G.D.; Dorigato, A.; Pegoretti,
A. Biogenic architectures for green, cheap, and efficient thermal energy storage and management. Renew. Energy 2021, 178, 96–
107. https://doi.org/10.1016/j.renene.2021.06.068.
113. Molinari, C.; Zanelli, C.; Laghi, L.; De Aloysio, G.; Santandrea, M.; Guarini, G.; Conte, S.; Dondi, M. Effect of scale-up on the
properties of PCM-impregnated tiles containing glass scraps. Case Stud. Constr. Mater. 2021, 14, e00526.
https://doi.org/10.1016/j.cscm.2021.e00526.
114. Rathore, P.K.S.; Kumar Shukla, S. Improvement in thermal properties of PCM/Expanded vermiculite/expanded graphite shape
stabilized composite PCM for building energy applications. Renew. Energy 2021, 176, 295–304.
https://doi.org/10.1016/j.renene.2021.05.068.
115. Yousefi, A.; Tang, W.; Khavarian, M.; Fang, C. Development of novel form-stable phase change material (PCM) composite using
recycled expanded glass for thermal energy storage in cementitious composite. Renew. Energy 2021, 175, 14–28.
https://doi.org/10.1016/j.renene.2021.04.123.
116. RAL Gütesicherung. Phase Change Materials RAL-GZ 896; RAL German Institute for Quality Assurance and Certification: Sankt
Augustin, Germany, 2018; p. 47.
117. ASTM C1784-20; Standard Test Method for Using a Heat Flow Meter Apparatus for Measuring Thermal Storage Properties of
Phase Change Materials and Products. ASTM International: West Conshohocken, PA, USA, 2020; Available online:
https://www.astm.org/Standards/C1784.htm (accessed on 25 September 2020).
118. International Energy Agency (IEA). Heat Pumps—Analysis; Tracking Report; International Energy Agency (IEA): Paris, France,
2020.
119. Vocale, P.; Morini, G.L.; Spiga, M. Influence of Outdoor Air Conditions on the Air Source Heat Pumps Performance. Energy
Procedia 2014, 45, 653–662. https://doi.org/10.1016/j.egypro.2014.01.070.
120. Javadi, H.; Mousavi Ajarostaghi, S.S.; Rosen, M.A.; Pourfallah, M. Performance of ground heat exchangers: A comprehensive
review of recent advances. Energy 2019, 178, 207–233. https://doi.org/10.1016/j.energy.2019.04.094.
121. You, T.; Wu, W.; Shi, W.; Wang, B.; Li, X. An overview of the problems and solutions of soil thermal imbalance of ground-
coupled heat pumps in cold regions. Appl. Energy 2016, 177, 515–536. https://doi.org/10.1016/j.apenergy.2016.05.115.
122. Wu, Y.; Gan, G.; Gonzalez, R.G.; Verhoef, A.; Vidale, P.L. Prediction of the thermal performance of horizontal-coupled ground-
source heat exchangers. Int. J. Low-Carbon Technol. 2011, 6, 261–269. https://doi.org/10.1093/ijlct/ctr013.
123. Habibi, M.; Hakkaki-Fard, A. Evaluation and improvement of the thermal performance of different types of horizontal ground
heat exchangers based on techno-economic analysis. Energy Convers. Manag. 2018, 171, 1177–1192. https://doi.org/10.1016/j.en-
conman.2018.06.070.
124. Dehdezi, P.K.; Hall, M.R.; Dawson, A.R. Enhancement of Soil Thermo-Physical Properties Using Microencapsulated Phase
Change Materials for Ground Source Heat Pump Applications. Appl. Mech. Mater. 2012, 110–116, 1191–1198.
https://doi.org/10.4028/www.scientific.net/AMM.110-116.1191.
125. Bottarelli, M.; Bortoloni, M.; Su, Y.; Yousif, C.; Aydın, A.A.; Georgiev, A. Numerical analysis of a novel ground heat exchanger
coupled with phase change materials. Appl. Therm. Eng. 2015, 88, 369–375. https://doi.org/10.1016/j.applthermaleng.2014.10.016.
126. Wang, J.L.; De Zhao, J.; Liu, N. Numerical Simulation of Borehole Heat Transfer with Phase Change Material as Grout. Appl.
Mech. Mater. 2014, 577, 44–47. https://doi.org/10.4028/www.scientific.net/AMM.577.44.
127. Chen, F.; Mao, J.; Chen, S.; Li, C.; Hou, P.; Liao, L. Efficiency analysis of utilizing phase change materials as grout for a vertical
U-tube heat exchanger coupled ground source heat pump system. Appl. Therm. Eng. 2018, 130, 698–709.
https://doi.org/10.1016/j.applthermaleng.2017.11.062.
128. Qi, D.; Pu, L.; Sun, F.; Li, Y. Numerical investigation on thermal performance of ground heat exchangers using phase change
materials as grout for ground source heat pump system. Appl. Therm. Eng. 2016, 106, 1023–1032. https://doi.org/10.1016/j.ap-
plthermaleng.2016.06.048.
129. Zhang, M.; Liu, X.; Biswas, K.; Warner, J. A three-dimensional numerical investigation of a novel shallow bore ground heat
exchanger integrated with phase change material. Appl. Therm. Eng. 2019, 162, 114297. https://doi.org/10.1016/j.ap-
plthermaleng.2019.114297.
130. Bonamente, E.; Aquino, A. Environmental Performance of Innovative Ground-Source Heat Pumps with PCM Energy Storage.
Energies 2020, 13, 117. https://doi.org/10.3390/en13010117.
131. Li, X.; Tong, C.; Duanmu, L.; Liu, L. Research on U-tube Heat Exchanger with Shape-stabilized Phase Change Backfill Material.
Procedia Eng. 2016, 146, 640–647. https://doi.org/10.1016/j.proeng.2016.06.420.
Polymers 2022, 14, 620 33 of 33
132. Yang, W.; Xu, R.; Yang, B.; Yang, J. Experimental and numerical investigations on the thermal performance of a borehole ground
heat exchanger with PCM backfill. Energy 2019, 174, 216–235. https://doi.org/10.1016/j.energy.2019.02.172.
133. Barbi, S.; Barbieri, F.; Marinelli, S.; Rimini, B.; Merchiori, S.; Larwa, B.; Bottarelli, M.; Montorsi, M. Phase change material-sand
mixtures for distributed latent heat thermal energy storage: Interaction and performance analysis. Renew. Energy 2021, 169,
1066–1076. https://doi.org/10.1016/j.renene.2021.01.088.
134. Zhu, N.; Hu, P.; Lei, Y.; Jiang, Z.; Lei, F. Numerical study on ground source heat pump integrated with phase change material
cooling storage system in office building. Appl. Therm. Eng. 2015, 87, 615–623. https://doi.org/10.1016/j.ap-
plthermaleng.2015.05.056.
135. Pu, L.; Xu, L.; Zhang, S.; Li, Y. Optimization of ground heat exchanger using microencapsulated phase change material slurry
based on tree-shaped structure. Appl. Energy 2019, 240, 860–869. https://doi.org/10.1016/j.apenergy.2019.02.088.
136. Kong, M.; Alvarado, J.L.; Thies, C.; Morefield, S.; Marsh, C.P. Field evaluation of microencapsulated phase change material
slurry in ground source heat pump systems. Energy 2017, 122, 691–700. https://doi.org/10.1016/j.energy.2016.12.092.
137. Bonamente, E.; Aquino, A.; Cotana, F. A PCM Thermal Storage for Ground-source Heat Pumps: Simulating the System Perfor-
mance via CFD Approach. Energy Procedia 2016, 101, 1079–1086. https://doi.org/10.1016/j.egypro.2016.11.147.
138. Bayomy, A.M.; Nguyen, H.V.; Wang, J.; Dworkin, S.B. Performance analysis of a single underground thermal storage borehole
using phase change material. In Proceedings of the International Ground Source Heat Pump Association (IGSHPA) Research
Track Conference, Stockholm, Sweden, 18–20 September 2018; pp. 1–11. https://doi.org/10.22488/okstate.18.000008.
139. Chen, F.; Mao, J.; Li, C.; Hou, P.; Li, Y.; Xing, Z.; Chen, S. Restoration performance and operation characteristics of a vertical U-
tube ground source heat pump system with phase change grouts under different running modes. Appl. Therm. Eng. 2018, 141,
467–482. https://doi.org/10.1016/j.applthermaleng.2018.06.009.
140. Faizal, M.; Bouazza, A.; Singh, R.M. Heat transfer enhancement of geothermal energy piles. Renew. Sustain. Energy Rev. 2016,
57, 16–33. https://doi.org/10.1016/j.rser.2015.12.065.
141. Mohamad, Z.; Fardoun, F.; Meftah, F. A review on energy piles design, evaluation, and optimization. J. Clean. Prod. 2021, 292,
125802. https://doi.org/10.1016/j.jclepro.2021.125802.
142. Han, C.; Yu, X. An innovative energy pile technology to expand the viability of geothermal bridge deck snow melting for dif-
ferent United States regions: Computational assisted feasibility analyses. Renew. Energy 2018, 123, 417–427.
https://doi.org/10.1016/j.renene.2018.02.044.
143. Mousa, M.M.; Bayomy, A.M.; Saghir, M.Z. Experimental and Numerical Study on Energy Piles with Phase Change Materials.
Energies 2020, 13, 4699. https://doi.org/10.3390/en13184699.
144. Sommerfeldt, N.; Madani, H. In-depth techno-economic analysis of PV/Thermal plus ground source heat pump systems for
multi-family houses in a heating dominated climate. Sol. Energy 2019, 190, 44–62. https://doi.org/10.1016/j.solener.2019.07.080.
145. Emmi, G.; Zarrella, A.; De Carli, M. A heat pump coupled with photovoltaic thermal hybrid solar collectors: A case study of a
multi-source energy system. Energy Convers. Manag. 2017, 151, 386–399. https://doi.org/10.1016/j.enconman.2017.08.077.