Fact-checked by Grok 2 weeks ago

Phase-change material

A phase-change material (PCM) is a substance that absorbs or releases a significant amount of during a , typically between solid and liquid states, enabling efficient and temperature regulation with minimal volume change. These materials are characterized by high energy storage density—often 5 to 14 times greater than storage methods—due to the isothermal nature of the phase change process, which occurs at a nearly constant temperature. PCMs are broadly classified into three categories: (e.g., paraffins and fatty acids, with latent heats of 150–250 kJ/kg and melting points from 20°C to 60°C), inorganic (e.g., salt hydrates, offering 170–330 kJ/kg and higher thermal conductivity of 0.5–1.0 W/m·K but prone to and ), and eutectic mixtures (combinations of the above for tailored temperatures and higher volumetric latent heat storage densities). Desirable properties include to prevent , , non-toxicity, and low flammability, though challenges like low in organics (0.1–0.3 W/m·K) often necessitate enhancements such as encapsulation or composite formulations with expanded or nanoparticles. The primary applications of PCMs leverage their ability to moderate temperature fluctuations and store , including passive building cooling (e.g., integration into walls or ceilings to reduce peak loads by up to 35%), systems (e.g., in Trombe walls for off-peak heat release), and battery management (e.g., preventing overheating in lithium-ion cells), and niche uses like or protective textiles for firefighters. Recent advances focus on shape-stabilized composites and to address leakage and improve heat transfer, expanding their role in solutions amid growing demands for .

Fundamentals

Definition and Principles

Phase-change materials (PCMs) are substances designed to store and release large amounts of through reversible transitions, typically solid-liquid or solid-solid, occurring at a nearly constant temperature. These materials exploit the high of inherent in the change process, allowing them to absorb when (endothermic) and release it when solidifying (exothermic) without significant temperature variation. This isothermal behavior makes PCMs particularly effective for thermal management where stable temperatures are required. The thermodynamic foundation of PCMs lies in storage, distinct from storage in materials like or , where is absorbed or released primarily through changes (Q = m * c * , with c as ). In contrast, for PCMs is quantified by the equation = m * L, where is the change, m is the , and L is the specific of , often ranging from 100 to 300 kJ/kg for common PCMs. This mechanism provides a volumetric density up to 10 times higher than methods at equivalent ranges, enabling compact and efficient systems. A representative example of a solid-liquid PCM is , which transitions from solid to liquid at temperatures around 50–60°C, storing approximately 200 kJ/kg of while maintaining phase stability over multiple cycles. The concept of PCMs gained prominence in the 1970s during the global energy crisis, when research focused on their use for storage to address intermittent supply challenges; the term "phase-change material" emerged in this context.

Phase Transition Mechanisms

Phase-change materials (PCMs) primarily utilize solid-liquid transitions for , as these involve substantial absorption or release during melting and freezing at a relatively constant . This process is governed by thermodynamic principles where the absorbs to overcome intermolecular forces, transitioning from an ordered solid lattice to a more disordered state without significant temperature change until the is complete. Solid-solid transitions, involving structural rearrangements within the solid phase (e.g., from one crystalline form to another), are less common due to their typically lower values and slower , though they avoid issues like leakage associated with . Liquid-gas transitions, while offering higher latent heats, are rarely employed in practical PCM applications because they entail large volume expansions (up to 1000 times) and require high pressures to contain the vapor phase, rendering them inefficient for compact storage systems. Solid-gas transitions, such as , are even more impractical owing to extreme volume changes and low energy densities under ambient conditions. The mechanisms underlying these phase transitions begin with nucleation, the initial formation of stable phase embryos, followed by growth where the new phase propagates through the material. In solid-liquid PCMs, heterogeneous nucleation often dominates during freezing, initiated at impurities, container walls, or additives that lower the energy barrier for crystal formation; homogeneous nucleation, occurring spontaneously in the bulk, requires significant undercooling and is less common. During melting, the reverse process involves dissolution of the solid lattice, with growth rates influenced by heat transfer rates and material purity. Supercooling, a key kinetic effect, occurs when the liquid phase persists below the equilibrium freezing temperature due to insufficient nucleation sites, potentially delaying energy release and reducing system efficiency; this can exceed 10-20% of the melting point in some materials. Hysteresis refers to the temperature offset between melting and freezing points, arising from supercooling, kinetic barriers, or structural changes, which can widen the effective transition range and impact thermal management precision. Efficiency in PCM phase transitions is often evaluated using figures of merit that balance latent heat storage against sensible heat losses over the transition temperature range. A key metric is defined as St = \frac{\Delta H}{\Delta T \cdot C_p}, where \Delta H is the latent heat of fusion, \Delta T is the temperature range of the transition (including hysteresis and supercooling effects), and C_p is the specific heat capacity; higher values indicate superior performance by maximizing latent heat relative to sensible contributions. Narrow transition ranges (\Delta T \approx 0-5^\circC) are ideal, as they minimize sensible heat dilution and enable isothermal operation, enhancing overall storage density. This metric underscores the preference for materials with sharp, reversible transitions to optimize energy retention in applications like building thermal regulation. A critical distinction in phase transition behavior is between congruent and incongruent melting, particularly relevant for inorganic PCMs like salt hydrates. Congruent melting occurs when the solid phase transforms directly into a of identical , ensuring reversibility and over cycles without . In contrast, incongruent melting involves into a and a solid phase of different (e.g., salt precipitating from a hydrated melt), leading to , reduced in subsequent cycles, and potential container . Salt hydrates such as decahydrate exemplify this issue, where incongruent behavior causes the solid phase to settle, necessitating additives or encapsulation to maintain uniformity. This mechanism directly affects long-term reliability, with congruent materials offering higher cycle .

Classification

Organic Phase-Change Materials

Organic phase-change materials (PCMs) constitute a primary class of carbon-based, non-polar compounds that primarily exhibit solid-liquid phase transitions for thermal energy storage. These materials are derived from natural or synthetic sources and include subtypes such as paraffins, fatty acids, polyalcohols, and polymers. Paraffins, consisting of linear alkanes like n-eicosane (C20H42), serve as representative examples with a melting point of approximately 36.9°C and a latent heat of fusion around 247 kJ/kg, making them suitable for applications near room temperature. Fatty acids, such as stearic acid (C18H36O2), offer higher melting points of about 69°C and latent heats of roughly 200 kJ/kg, providing versatility for moderate-temperature storage. Polyalcohols, including fatty alcohols like octadecanol, and polymers such as polyethylene glycol (PEG), further expand the category; PEG, for instance, allows tunable phase transition temperatures through molecular weight variations while maintaining high enthalpies of fusion exceeding 150 kJ/kg. These organic PCMs demonstrate several key advantages that enhance their practicality in thermal management systems. They exhibit excellent over repeated cycles, with no or significant , ensuring consistent performance. Unlike inorganic counterparts, organic PCMs are non-corrosive to common construction materials, possess low , and undergo , which preserves their composition during transitions. Additionally, they cover a broad range from -5°C to 120°C, accommodating diverse applications without the need for complex handling. Despite these benefits, organic PCMs face notable limitations that can impact efficiency. Their thermal conductivity is inherently low, typically ranging from 0.1 to 0.2 W/m·K for paraffins and fatty acids, which slows rates and requires enhancements for practical use. Flammability poses a concern, particularly for paraffins, necessitating fire-retardant measures in enclosed systems. Furthermore, phase transitions involve volume changes of 10-20%, leading to potential mechanical stress or leakage if not properly contained. To address sustainability challenges, bio-based PCMs derived from oils and waste fats have emerged prominently since the , offering renewable alternatives to petroleum-derived options. These materials, such as those processed from or soy oils into esters, maintain comparable thermophysical properties while reducing environmental impact through biodegradable sourcing. Research highlights their potential for eco-friendly thermal storage, with examples achieving melting points around 30-40°C and latent heats over 180 kJ/kg.

Inorganic Phase-Change Materials

Inorganic phase-change materials (PCMs) encompass a diverse of substances that undergo solid-liquid transitions, offering high capacities suitable for thermal management applications. These materials are primarily categorized into hydrates, pure s, and metals or alloys, each exhibiting distinct thermophysical that make them advantageous for specific ranges. Unlike PCMs, inorganics generally provide superior volumetric due to their higher densities and latent heats, though they present unique challenges. Salt hydrates, such as decahydrate (Na₂SO₄·10H₂O), represent a prominent subtype with temperatures around room conditions, for instance, melting at 32.4°C and delivering a of 239–254 kJ/kg. Pure salts like (NaNO₃) target higher temperatures, with a of 306.4°C and of 178.6 kJ/kg, making them viable for industrial recovery. Metals and alloys, exemplified by , operate at low temperatures ( 29.8°C, 80.1 kJ/kg) but compensate with high (approximately 5.9 g/cm³), yielding a volumetric storage of about 488 kJ/L. These subtypes collectively enable applications from building cooling to thermal regulation, leveraging their inherent high s up to 300 kJ/kg and thermal conductivities in the range of 0.5–1 W/m·K, which facilitate faster compared to organics. Additionally, their low cost (often 1–20 $/kWh thermal capacity) and non-flammable nature enhance safety and economic viability. Despite these benefits, inorganic PCMs face significant limitations that can impair long-term performance. Salt hydrates are particularly prone to phase segregation during incongruent melting, where the anhydrous salt separates from the water, resulting in a progressive loss of latent heat over repeated cycles—up to 20–30% degradation after 100 cycles in untreated materials. Supercooling, another common issue, can delay solidification by 10–20°C below the melting point, reducing the effective operating temperature range and energy recovery efficiency. Furthermore, many inorganics, especially chlorides and nitrates, exhibit corrosiveness toward common container metals like steel, with corrosion rates reaching 70 mg/cm² after extended cycling, necessitating protective coatings or compatible materials. Eutectic salt mixtures, such as NaCl-Na₂CO₃, address some stability concerns for high-temperature applications above 200°C, like concentrated solar thermal power, maintaining a latent heat of 311.6 kJ/kg at 635°C over 1000 cycles with minimal property degradation.

Solid-Solid and Eutectic Phase-Change Materials

Solid-solid phase-change materials (PCMs) represent a class of materials that undergo a between two solid states, typically from crystalline to amorphous, without passing through a . This transition allows them to store and release while maintaining structural integrity, making them particularly suitable for applications where leakage is a concern. Polymeric materials, such as , are commonly used examples, exhibiting phase transitions in the range of 40-60°C with latent heats of 100-200 kJ/kg. The primary advantages of solid-solid PCMs include the absence of leakage during phase change, as no liquid intermediate forms, and minimal volume change, typically less than 1%, which enhances compatibility with surrounding structures. These properties enable their use without encapsulation in certain scenarios, simplifying integration into composites or coatings. However, limitations such as relatively lower compared to solid-liquid PCMs, more complex synthesis processes involving crosslinking, and higher production costs can restrict widespread adoption. Eutectic phase-change materials, on the other hand, are mixtures of two or more components—often and inorganic substances—that exhibit a sharp at a specific , behaving as a single during . These mixtures form at the eutectic point in phase diagrams, where the liquidus lines of the constituent components intersect, resulting in a lowest possible for the system. A representative example is the eutectic mixture of and , which transitions at approximately 20°C with a of 180 kJ/kg, making it suitable for low-temperature thermal management.

Properties and Selection

Thermophysical Properties

Phase-change materials (PCMs) exhibit several key thermophysical properties that determine their efficacy in . The melting or freezing (Tm) typically ranges from below 0°C to over 100°C, depending on the material type, enabling selection for diverse regimes. The of fusion (ΔHf) is a primary attribute, generally falling between 100 and 250 kJ/kg for most and inorganic PCMs, such as 128–244 kJ/kg for paraffins and 105–231 kJ/kg for salt hydrates. (Cp) varies from 1 to 3 kJ/kg·K in both solid and liquid phases, influencing storage. Thermal conductivity (k) is often low, ranging from 0.1 to 0.5 W/m·K for pure PCMs and up to 1 W/m·K for inorganics, limiting rates. (ρ) spans 700–1600 kg/m³, with organics around 800–900 kg/m³ and inorganics higher at 1400–1600 kg/m³. Volume expansion during phase change is notable, typically 5–20%, with minimal changes (around 5–10%) for organics like paraffins and larger (10–20%) for inorganics, necessitating container design considerations.
PropertyTypical RangeExample Values (Organic PCMs)
Melting Temperature (Tm)-5°C to 200°C45–55°C (paraffins)
Latent Heat (ΔHf)100–300 kJ/kg160–170 kJ/kg (RT series)
Specific Heat (Cp)1–3 kJ/kg·K2 kJ/kg·K (solid/liquid)
Thermal Conductivity (k)0.1–1 W/m·K0.2 W/m·K (paraffins)
Density (ρ)700–1600 kg/m³770–880 kg/m³ (liquid/solid)
Volume Expansion5–20%10% (paraffins)
These ranges highlight differences between and inorganic PCMs, with organics generally showing lower thermal conductivity but better . measurement techniques ensure accurate characterization of these properties. (DSC) is widely used to determine Tm and ΔHf, involving heating or cooling small samples (5–10 mg) at controlled rates (e.g., 5–10°C/min) under ASTM E793 standards, providing precise endothermic/exothermic peaks for phase transitions. Thermal gravimetric analysis (TGA) assesses thermal stability by monitoring mass loss with temperature, revealing decomposition onset above 200–300°C for most PCMs. The hot disk method, a transient plane source technique, measures k by analyzing temperature responses to a heated disk embedded in the sample, suitable for both and phases. These methods allow for reliable empirical data, though DSC is limited to small samples and may vary with scan rates. Property trade-offs in PCMs emphasize the dominance of latent heat in total energy storage capacity. The overall heat stored (Q) per unit mass is given by Q = \Delta H_f + \int_{T_1}^{T_2} C_p \, dT where the latent heat term (ΔHf) typically contributes 70–90% of Q during phase transition at Tm, far outweighing sensible heat from Cp over temperature changes (ΔT ≈ 10–20 K), underscoring why high ΔHf is prioritized despite trade-offs like low k that slow charging/discharging. Cycle is evaluated through repeated melting/freezing tests, with many PCMs demonstrating retention of over 95% of initial ΔHf after 1000 or more cycles, indicating minimal from or phase segregation; for instance, certain paraffin-based PCMs retain 99.5% after 50 cycles and remain stable beyond 1000 under controlled conditions.

Selection Criteria

The selection of phase-change materials (PCMs) for specific applications involves evaluating key criteria to ensure optimal performance, reliability, and economic viability. The desired temperature range is a primary consideration, such as 20–30°C for comfort cooling systems, to align with operational demands and maximize or release . , expressed as per unit volume, determines the storage capacity relative to system size, prioritizing materials that provide high volumetric changes without excessive bulk. cycle life is critical for long-term durability, with suitable PCMs demonstrating over 1000–5000 cycles to minimize and needs. Cost-effectiveness varies by type, typically $2–5/kg for basic organic PCMs and up to $8/kg for encapsulated or bio-based variants as of 2025, balancing initial investment against lifecycle benefits for applications like building integration. compatibility ensures non-corrosive behavior with systems, preventing structural damage or leakage during repeated . To facilitate comparative analysis, figures of merit quantify trade-offs among properties. The energy storage index (ESI), defined as \text{ESI} = \frac{\Delta H_f}{\Delta T}, where \Delta H_f is the of fusion and \Delta T is the temperature range, evaluates storage efficiency by normalizing energy capacity against the width of the melting/freezing interval. The conductivity factor, k / \Delta T, where k is , highlights materials that enable rapid within narrow temperature spans, addressing limitations like the inherently low k of PCMs. An overall suitability score integrates multiple attributes—such as , , , cost, and stability—using multi-criteria decision methods like to rank PCMs for targeted uses. Environmental factors further guide selection toward sustainable options. Toxicity assessments prioritize low-hazard materials to ensure in enclosed or human-contact applications, while biodegradability favors bio-based PCMs that decompose without persistent pollutants. emphasizes renewable sourcing, such as plant-derived over petroleum-based PCMs, to reduce carbon footprints and align with principles. Recent composites, as of 2025, show enhanced cycle stability exceeding 10,000 cycles in some cases, supporting broader adoption in energy-efficient systems.

Development and Technology

Historical Development

The concept of , fundamental to phase-change materials (PCMs), was first systematically described by Scottish physicist in the 1760s through his experiments on the heat absorbed during the melting of ice without a temperature rise. Practical applications of PCMs emerged in the mid-20th century, notably in 1948 when biophysicist Maria Telkes developed the world's first solar-heated house in , utilizing Glauber's salt (sodium sulfate decahydrate) as a PCM to store from passive collection. This innovation marked an early shift toward using PCMs for (TES), though initial efforts were limited by material stability issues like phase segregation in inorganic salts. The 1970s oil crisis catalyzed widespread research into energy-efficient technologies, prompting the U.S. Department of Energy (DoE) and to investigate PCMs for TES applications. 's programs, initiated around 1970, integrated PCMs into space missions for thermal control, such as in habitats and the , where materials like paraffin waxes provided reliable heat management in extreme environments. These efforts, driven by the need to address energy supply mismatches, expanded PCM use from passive solar systems—exemplified by Telkes' work—to active TES configurations, including solar thermal storage tanks. By the late 1970s, DoE-funded studies emphasized PCMs, such as fatty acids, for their , laying groundwork for broader commercialization. The 1980s saw the advent of the first commercial paraffin-based PCMs, primarily for storage, with products like encapsulated wax modules enabling efficient retention in residential and industrial systems. This period reflected a transition toward scalable TES, spurred by ongoing demands, and included early military adaptations, such as PCM-integrated vests for personnel temperature regulation in harsh conditions. In the , encapsulation technologies advanced significantly, with patents for microcapsules—such as those developed by for polymer-shelled cores—addressing leakage and enhancing durability for building integration. These innovations, including melamine-formaldehyde shells, facilitated the incorporation of PCMs into boards and textiles, boosting efficiency. The early 2000s intensified focus on PCMs in building applications following the 1997 , which emphasized reductions and prompted research into and heating systems to minimize reliance. This era highlighted a pivot from solar-centric uses to comprehensive TES in architecture, with organic PCMs like paraffins briefly referenced for their non-corrosive properties in such contexts. By the , a boom in nano-enhanced PCMs emerged, incorporating nanoparticles like or carbon nanotubes to improve thermal conductivity, resulting in over 800 publications by 2021 documenting these advancements for high-impact applications in .

Encapsulation Methods

Encapsulation methods are crucial for phase-change materials (PCMs) to mitigate issues such as leakage during , chemical reactivity with surrounding media, and volume changes that could compromise system integrity. By enclosing the PCM core in a protective shell, these techniques form stable core-shell structures that enable integration into matrices like building materials or fluids without compromising the PCM's storage capacity. Encapsulation also enhances handling, prevents in some cases, and improves compatibility, making PCMs viable for practical applications. Encapsulation is categorized by scale into macroencapsulation, , and nanoencapsulation. Macroencapsulation involves enclosing bulk PCMs in sealed containers, such as (HDPE) spheres or tubes with diameters ranging from 1 to 50 mm, which are suitable for large-scale systems like solar thermal storage where ease of replacement and minimal processing are prioritized. creates smaller particles, typically 1 to 1000 μm in diameter, using polymer shells to contain the PCM core; for instance, resins are commonly employed to encapsulate paraffin-based PCMs, preventing leakage and allowing dispersion in composites like boards. Nanoencapsulation further reduces the size to below 100 nm, often by impregnating PCMs into nanostructures such as carbon nanotubes or mesoporous silica, which boosts surface area and thermal response for advanced applications like fluids. Key fabrication methods for micro- and nanoencapsulation include in-situ polymerization, emulsion-based techniques, and . In in-situ polymerization, the shell material polymerizes directly around emulsified PCM droplets, forming a robust barrier; this method is widely used for producing melamine-formaldehyde or microcapsules with high encapsulation efficiency. involves dispersing the PCM in an oil-in-water , where monomers in the continuous phase react to coat the droplets, offering control over shell thickness and uniformity for paraffins or hydrates. atomizes a PCM-shell into a hot gas stream to evaporate the and solidify microcapsules, a scalable process ideal for heat-sensitive PCMs that yields spherical particles with reduced agglomeration. These methods address critical challenges in PCM deployment. Core-shell structures significantly reduce leakage, with well-designed microcapsules maintaining integrity over thousands of thermal cycles by containing the liquid phase within impermeable barriers. For corrosive inorganic PCMs like salt hydrates, silica-based shells provide chemical inertness and prevent reactions with external environments, enhancing long-term stability. Encapsulation can also mitigate the inherently low thermal conductivity of PCMs by incorporating conductive shell materials, though this is secondary to goals. A notable commercial example is BASF's Micronal, introduced in the early , which uses melamine-formaldehyde shells to microencapsulate waxes for building applications; recent advancements in the field have shifted toward bio-based shells to improve and reduce environmental impact.

Advanced Materials

Thermal Composites

Thermal composites integrate phase-change materials (PCMs) with supportive matrices, such as porous or polymeric structures, to enhance overall thermal performance while maintaining storage capabilities. Unlike pure PCMs, which often suffer from low thermal conductivity—exemplified by typical value of approximately 0.2 W/m·K—these hybrids address limitations through structural and conductive additives. This integration enables better efficiency in applications requiring rapid charging and discharging. Key types include shape-stabilized PCMs, where liquid PCMs are confined within a porous to prevent leakage during phase transitions. For instance, impregnated into expanded (EG) forms a composite that boosts thermal conductivity to 1–10 W/m·K, depending on EG loading (e.g., up to 10.70 W/m·K at 10 wt.% EG via prefabricated skeleton methods), representing about a 20-fold increase over pure while retaining over 87% of . Form-stable composites, another prominent category, employ polymer matrices like (HDPE) to encapsulate PCMs at loadings exceeding 80 wt.%, ensuring structural integrity without exudation even under repeated cycling. Enhancement mechanisms primarily focus on improving thermal conductivity through the addition of high-conductivity fillers, such as or metal foams, which create efficient heat conduction pathways within the matrix. , for example, can increase effective conductivity in fiber/paraffin composites by aligning to form networks that facilitate directional heat flow. Metal foams provide a sturdy, porous that significantly elevates rates in PCMs by up to several times the base value, owing to their high and metallic ligaments. A simplified model for estimating effective thermal conductivity in such systems is the approximation: k_{\text{eff}} = k_{\text{matrix}} + \phi (k_{\text{PCM}} - k_{\text{matrix}}) where k_{\text{eff}} is the effective thermal conductivity, k_{\text{matrix}} and k_{\text{PCM}} are the conductivities of the matrix and PCM, respectively, and \phi is the volume fraction of the PCM; this linear relation provides a foundational estimate for composite design, though more advanced models account for interfacial effects. These composites offer distinct advantages, including leak-proof operation due to confinement or binding, the ability to mold into desired shapes for versatile integration, and enhanced long-term durability with cycle stabilities reaching up to 10,000 melt-freeze cycles, showing minimal degradation in (e.g., <10% loss). Since 2015, -based PCM composites have emerged as a specialized variant, leveraging anisotropic hybrid scaffolds (e.g., with reduced oxide and expanded ) to achieve thermal conductivities up to 0.79 W/m·K in aligned (longitudinal) directions, making them promising for where , high-performance materials are critical. Recent 2023-2025 studies have further enhanced anisotropic in PCMs, achieving up to 4.36 W/m·K through-plane conductivity with low loading (1.07 vol.%), improving applications in solar- .

Photo-Thermal Conversion Composites

Photo-thermal conversion composites integrate phase-change materials (PCMs) with photothermal absorbers to enable efficient capture and storage of through light-to-heat conversion followed by storage. Typical compositions feature PCMs like combined with broadband absorbers such as , , or plasmonic nanoparticles, which exhibit high absorptivity in the ultraviolet-visible-near-infrared spectrum. For example, -graphene oxide composites leverage the high surface area and conductivity of graphene oxide to achieve solar-thermal conversion efficiencies exceeding 90%. These materials address limitations in pure PCMs by enhancing light absorption while maintaining phase-change stability. The core mechanism involves photothermal conversion, where incident solar radiation is absorbed by the photothermal component and rapidly transformed into heat, quantified by the efficiency formula \eta = \frac{h A \Delta T}{I \alpha}, where h is the , A the surface area, \Delta T the rise, I the , and \alpha the absorptivity. This generated heat elevates the composite to the PCM's , enabling via without significant leakage, due to the structural support from the absorbers. The process ensures high temporal matching between solar input and thermal output. Key advantages of these composites include direct solar-to-thermal , which reduces transmission losses compared to indirect methods, and minimized radiative dissipation through full-spectrum utilization. MXene-based PCM composites exemplify this, delivering photothermal efficiencies over 85%—with optimized formulations reaching 94.5% under one-sun illumination—while providing form-stability and high . Photo-thermal conversion phase-change composite materials (PTCPCESMs), emerging post-2018, further advance these benefits; 2023 investigations into MXene-integrated variants demonstrated 20% higher yields relative to non-composite systems, attributed to enhanced rates up to 92.6% efficiency.

Applications

Building and Construction

Phase-change materials (PCMs) are integrated into building envelopes to provide passive thermal regulation, primarily by absorbing and releasing during phase transitions to stabilize indoor temperatures and reduce reliance on (HVAC) systems. Common integration methods include incorporating microencapsulated organic PCMs, such as paraffins with phase transition temperatures of 20-25°C, into wallboards like panels. These microcapsules prevent leakage during melting and allow for direct mixing during manufacturing, enhancing the material's without significantly altering its structural properties. Similarly, PCMs can be added to as aggregates or admixtures, where they impregnate porous structures to increase the building's capacity while maintaining integrity. Another approach involves phase-change windows, where PCM layers are in glazing units or frames to mitigate gain and improve , particularly in double-glazed systems. The primary benefits of PCM integration in include peak load shifting, where stored is released during off-peak hours to balance demand, potentially reducing overall energy consumption by up to 30%. This enhances thermal inertia, delaying through walls and roofs, which improves occupant comfort and cuts HVAC operational costs. For instance, in a of PCM-enhanced walls in hot climates like those in , cooling needs were reduced by 25% compared to conventional brick walls, demonstrating effective temperature moderation during peak daytime heat. These advantages are quantified through metrics such as daily heat storage capacity, typically ranging from 100-200 kJ/m² for wallboard applications, which supports diurnal thermal cycling. Performance is evaluated using standards like ISO 13786, which defines dynamic thermal properties including thermal admittance and decrement factor to assess how PCMs influence heat flow over time. Commercial products, such as BioPCM introduced in the , exemplify practical adoption, featuring bio-based microencapsulated PCMs suitable for ceilings, walls, and floors to achieve . Union-funded projects, including demonstrations under initiatives, have reported annual savings of 15-20% in building use by 2020 through PCM retrofits, validating their role in meeting sustainability targets.

Electronics and Thermal Management

Phase-change materials (PCMs) play a critical role in thermal management for electronics by absorbing transient heat loads during high-power operations, such as in portable devices and high-density circuits, where traditional convective cooling struggles with space constraints and noise. In compact systems like smartphones and laptops, PCM-integrated heat sinks utilize the latent heat of fusion to buffer temperature spikes, maintaining operational temperatures below critical thresholds without active fans. For instance, paraffin-based PCMs like Rubitherm RT42, with a melting point of 38–42°C and latent heat capacity of approximately 165 kJ/kg, have been employed in heat sink designs to absorb excess heat from processors during intensive tasks. In (EV) battery packs, PCMs are integrated to prevent by rapidly absorbing heat generated during fast charging or abuse conditions, thereby delaying propagation to adjacent cells. Salt hydrate PCMs, such as trihydrate composites, offer high storage (around 200–250 kJ/kg) and phase transitions near 50–60°C, which align with safe operating limits, significantly mitigating temperature rises and enhancing pack safety during runaway events. These materials are encapsulated in modules surrounding cylindrical or prismatic cells to provide uniform cooling without electrical interference. High cycle stability is essential for repeated charge-discharge cycles in EVs, ensuring minimal degradation over thousands of operations. Hybrid systems combining PCMs with s address the low thermal conductivity of pure PCMs (typically 0.2–0.5 W/m·K), enhancing heat spreading in cooling. These designs feature PCMs with change temperatures of 40–60°C to keep device surfaces under 50°C, where s transport absorbed to external sinks, improving overall efficiency by 20–30% compared to PCM alone. For example, in thermal management, such hybrids maintain cell temperatures below 45°C under 2C discharge rates, with the PCM absorbing peak loads while the ensures rapid dissipation. Performance metrics demonstrate that PCM-based systems can extend safe operation during overloads by 2–3 times relative to air-cooled alternatives, as the buffers heat until fully melted, delaying overheating. The cooling duration t can be estimated using the relation t = \frac{Q_{\text{load}}}{\Delta H_f \cdot m_{\text{PCM}}} where Q_{\text{load}} is the heat input, \Delta H_f is the of , and m_{\text{PCM}} is the PCM ; this simplifies transient for . In the 2020s, PCM adoption has grown in base stations and centers, where integrated heat sinks reduce reliance on fans, cutting cooling power consumption by up to 40% through passive storage during peak loads.

Emerging Industrial Uses

Phase-change materials (PCMs) are increasingly integrated into textiles through techniques, enabling temperature-regulating fabrics that enhance wearer comfort in varying environments. Technologies, originating from NASA-developed microencapsulated PCMs known as Thermocules, incorporates these into fibers and coatings for apparel such as jackets and sleepwear, where the PCMs absorb excess heat during activity and release it when cooling occurs. The materials typically transition at around 28°C, aligning with temperature, and provide storage in the range of 50-100 kJ/kg, allowing sustained thermal balance without active power. This application has been commercialized in products from brands like Burton and , demonstrating durability in outdoor and performance gear. In transportation, particularly , PCM panels are employed in reefer trucks to maintain precise ranges for sensitive goods like , reducing reliance on constant refrigeration and fuel consumption. These panels, often using organic PCMs such as decyl alcohol or n-tetradecane, phase change within the 2-8°C window critical for stability, absorbing heat during door openings or transit fluctuations to prevent spoilage. Studies on PCM-integrated refrigerated trucks show they can extend holding times by several hours compared to traditional systems, improving efficiency in . For instance, composite PCM materials have been tested to sustain potency during multi-day transports. High-temperature PCMs, primarily inorganic salts, are emerging in (CSP) systems to enable efficient beyond sunset, supporting continuous power generation. Salts like (melting point ~800°C, latent heat ~450 kJ/kg) and magnesium chloride (~714°C, ~452 kJ/kg) operate above 300°C, storing heat at higher densities than sensible storage methods and reducing overall system costs. Encapsulated designs mitigate corrosion issues at these temperatures, with pilots demonstrating viability in cascaded configurations for multi-hour dispatchability. In biomedical applications, PCMs facilitate controlled through implantable devices, where phase transitions trigger precise release profiles in response to physiological cues. Subcutaneous implants incorporating PCM-based micropumps, such as those using thermo-responsive materials, enable on-demand dosing for conditions by changing from to states at body , ensuring sustained therapeutic levels over weeks. Constructs like PCM spheres or blocks have been developed to regulate release of agents like , promoting targeted outcomes such as neurite outgrowth without burst effects. PCMs are also advancing grid-scale via phase-change thermal batteries, which capture excess as for later dispatch. In 2022 pilots, these systems integrated with absorption cycles showed up to 20% efficiency improvements over conventional thermal storage by optimizing heat recovery during off-peak periods. Such applications leverage high capacities to stabilize grids, with materials like salt hydrates enabling scalable, low-cost solutions for balancing intermittent and inputs.

Challenges and Safety

Fire and Stability Issues

Organic phase-change materials (PCMs), particularly paraffins, pose significant fire risks due to their flammability, with flash points typically around 200°C. These materials exhibit high heat release rates during , often reaching peak values of 800–1100 kW/m² as measured by cone tests under standard fluxes of 35–50 kW/m². Such behavior can accelerate fire spread in applications like building envelopes, where organic PCMs are integrated into walls or ceilings. Cone , a key testing method per ISO 5660, quantifies these risks by simulating real-fire conditions and evaluating parameters like ignition time and total release. Stability issues further complicate the long-term use of PCMs, as repeated melting-freezing cycles lead to chemical decomposition, including oxidation in air exposure, which degrades latent heat capacity by 10–20% after 1000 cycles in some organic formulations. Bio-based PCMs, derived from fatty acids or plant oils, are particularly vulnerable to microbial growth, which can cause structural breakdown and reduced thermal performance over time. These degradation mechanisms arise from environmental interactions, such as humidity and temperature fluctuations, compromising the material's integrity in practical settings. Mitigation strategies for fire risks include incorporating flame-retardant additives, such as coatings, which expand under to form a barrier, reducing peak release rates by approximately 50–60% in treated paraffin composites. For instance, adding ammonium polyphosphate or melamine-based retardants to PCMs has demonstrated substantial suppression of in cone tests. To address stability concerns, storing or operating PCMs in inert atmospheres like prevents oxidation, preserving thermophysical properties over extended cycles. Fire tests in the , including building simulations, underscored these vulnerabilities in PCMs, prompting a shift toward non-flammable hydrated alternatives for enhanced in . Standards like NFPA 68 guide venting designs to manage potential pressure buildup from PCM-related fires in enclosed building systems.

Environmental and Economic Considerations

Phase-change materials (PCMs) present varied environmental profiles depending on their composition. Inorganic PCMs, such as certain salt hydrates, can exhibit due to potential corrosivity and the presence of in some formulations, posing risks during handling and leakage. In contrast, organic PCMs, particularly bio-based variants like (), demonstrate high recyclability, with recovery rates exceeding 90% through processes like intermolecular interactions in form-stable composites. Bio-based PCMs, derived from renewable sources such as s or waste oils, generally have a lower than synthetic petroleum-derived options like , with life-cycle assessments showing reductions in of up to 40% for waste-derived PCMs compared to . Economically, PCM production costs range from approximately $2 to $5 per kilogram for raw materials and basic formulations, though encapsulation can increase this to $5-10 per kilogram, making scalability dependent on material choice and application. In building applications, PCM integration often yields payback periods of 3-7 years through energy savings, as evidenced by analyses of prefabricated structures where environmental and economic returns align within this timeframe. The global PCM market reached approximately $3 billion in 2025, driven by demand in energy storage and thermal management sectors. Key challenges include supply chain vulnerabilities for specialized inorganic PCMs relying on rare materials, such as salts used in some storage formulations, which face sourcing constraints and price volatility. End-of-life disposal is regulated under the REACH framework, which mandates risk assessments and safe management of chemical substances like PCMs to protect human health and the , requiring licensed facilities for potentially hazardous variants. (LCA) studies indicate that PCM-integrated buildings can reduce CO2 emissions by 20-30% over a 50-year lifespan through lowered operational demands, though embodied impacts from must be minimized for net benefits.

Recent Advances

Innovations in Composites and Encapsulation

Recent advancements in phase-change materials (PCMs) have focused on nano-composites to enhance thermal storage capacity and conductivity. Integration of metal-organic frameworks (MOFs) into PCMs has enabled significant improvements in storage. For instance, hierarchical rGO@MOF-5 composites achieve a of 168.7 J/g, marking an 18.5% increase over conventional counterparts, attributed to the porous structure of MOFs that facilitates better PCM impregnation and . Similarly, 2024 studies on aerogels as scaffolds in paraffin-based PCMs have demonstrated exceptional thermal conductivity enhancements, with hyperbolic aerogel composites reaching up to 82.4 W/m·K at 30 wt% loading, enabling efficient heat dissipation while maintaining shape stability through capillary confinement. Smart encapsulation techniques have emerged as a key innovation to mitigate leakage and improve durability in PCMs. Self-healing shells, incorporating matrices with embedded microcapsules, allow for autonomous repair of cracks or leaks during thermal cycling; for example, sustainable self-healing PCMs based on fatty alcohols exhibit repeated recovery of phase-change functionality after mechanical damage, extending operational lifespan in dynamic environments. In specialized applications, phase-change microcapsules have been developed for oilfield , where 2025 highlights their role in cooling drilling fluids by 2-12°C through absorption, using cores like encapsulated in melamine-urea-formaldehyde shells via in-situ , thus enhancing thermal management in high-temperature reservoirs. Oriented composites represent a notable innovation, where aligned fillers such as carbon fibers or graphene sheets direct thermal conductivity anisotropically, optimizing heat flow in targeted directions. Recent 2023-2025 studies on oriented phase-change composites (OCPCMs) have achieved exceptional directional heat transfer, with aligned carbon-based fillers enabling up to several-fold increases in in-plane conductivity compared to isotropic designs, addressing limitations in uniform heat spreading for advanced thermal systems. Encapsulated PCMs (ePCMs) have seen parallel progress in thermal management, with nano-enhanced variants incorporating SiO₂ or graphene shells yielding up to 97% conductivity boosts and 38% higher heat rejection in integrated heat sinks, as detailed in 2025 reviews. A 2025 comprehensive review on phase-change energy storage materials (PCESMs) underscores advances in nano-enhanced PCMs for thermal management applications.

New Applications and Research Directions

Recent research has explored the integration of phase-change materials (PCMs) in wearable biomedical devices, particularly for controlled drug delivery systems. For instance, heat-responsive phase-change creams have been developed to enhance transdermal insulin delivery, achieving up to 19.1% efficiency in skin appendages under photothermal stimulation, which supports more effective management of diabetes through targeted release at body temperature. In space exploration, NASA has advanced PCM-based thermal control systems for habitats, including Mars missions, where PCM panels provide passive energy storage to maintain stable temperatures in extreme environments, as detailed in 2024 state-of-the-art reports on small spacecraft technology. These systems absorb and release heat to protect electronics and crew quarters, with laboratory experiments demonstrating effective passive cooling for simulated space habitats. Emerging industrial applications include PCM microcapsules for oilfield operations, where they absorb excess heat during to prevent and improve in ultra-deep wells. A review highlights their ability to release controllably, reducing operational temperatures by up to 12°C and extending equipment lifespan in harsh conditions. In integration, PCMs enhance performance by mitigating overheating; for example, their incorporation into photovoltaic systems stabilizes temperatures, with some systems achieving up to 15.4% . A 2024 MDPI review on PCMs in buildings further underscores savings, reporting up to 30% reductions in heating and cooling demands through optimized integration, addressing gaps in urban . Research directions emphasize AI-driven optimization for PCM design, where machine learning models predict thermophysical properties like melting points and latent heat with high accuracy, accelerating material development for tailored applications. Sustainable sourcing from waste materials, such as bio-based organics, has shown promise in 2025 studies, yielding approximately 40% reductions in while maintaining performance in thermal storage. For electronic cooling, 2025 reviews advocate hybrid PCM systems that extend device lifespans by 50% under high loads, with ongoing trials in data centers. In electric vehicles, oriented composite PCMs for thermal management, advanced in 2025, improve charging rates and safety by uniformly distributing heat, potentially increasing range by 15%.

References

  1. [1]
    (PDF) A review on phase change materials and their applications
    Aug 10, 2025 · In phase change materials, defined as the materials that can store latent heat and sensible heat, there is a lesser storage in sensible heat, as ...
  2. [2]
    A Comprehensive Review on Phase Change Materials and ...
    Phase change materials (PCMs) have shown their big potential in many thermal applications with a tendency for further expansion. One of the application ...
  3. [3]
    Understanding Phase Change Materials (PCM) - Thermtest
    Dec 13, 2024 · Phase change materials (or PCMs) are materials that absorb and release large amounts of energy when they change phases, for example from solid to liquid or ...
  4. [4]
    Phase Change Materials for Life Science Applications
    Jan 25, 2023 · PCMs are a category of advanced materials that are used to store thermal energy as latent heat (Figure 2). They play an essential role in the ...Introduction · Phase Change Materials (PCMs) · Applications of PCMs in Life...
  5. [5]
    Review on thermal energy storage with phase change
    In this work, a review has been carried out of the history of thermal energy storage with solid–liquid phase change. Three aspects have been the focus of this ...Missing: patents | Show results with:patents
  6. [6]
    What are Phase Change Materials? (Will they be the next big thing ...
    Apr 14, 2025 · PCMs are materials that, when they change phase (melt or freeze), absorb or release heat at a nearly constant temperature, and by doing this can be effectively ...
  7. [7]
    Phase Change Material - an overview | ScienceDirect Topics
    In addition to solid–liquid type, the phase transition type in PCMs can be solid–solid as well, which occurs between amorphous and crystalline phases or two ...
  8. [8]
    Research progress in nucleation and supercooling induced by ...
    This paper reviews the research progress of controlling the supercooling and crystal nucleation of phase change materials.
  9. [9]
    Supercooling of phase change materials: A review - ScienceDirect
    Supercooling is a natural phenomenon that keeps a phase change material (PCM) in its liquid state at a temperature lower than its solidification temperature.
  10. [10]
    Suppression of supercooling and phase change hysteresis of Al ...
    Jul 30, 2024 · This study to investigate the effect of Ti-doping on the PCH and supercooling of Al-25mass%Si MEPCMs for efficient thermal energy storage and utilization.
  11. [11]
    Review Article Salt hydrate bio-composites for thermal energy storage
    Congruent melting of hygroscopic salts occurs when the salt is soluble in its hydrated water at melting whereas incongruent melting occurs when the salt is not ...
  12. [12]
    [PDF] Review of Inorganic Salt Hydrates with Phase Change Temperature ...
    Jul 9, 2018 · The salt's insolubility in the stoichiometric water of its hydrate causes incongruent melting, where anhydrous salt settles out of solution and ...
  13. [13]
    Thermo-physical properties of n-eicosane - ResearchGate
    The physical properties of PCM are: Melting point (36.9°C), Density (kg/m 3 ): [856 (solid), 780 (liquid)], [18] . ... ... Phase change materials (PCMs) are used ...
  14. [14]
    Numerical Study of N-Eicosane Melting Inside a Horizontal Cylinder ...
    -1, thermal conductivity 60 W m-1 K-1, latent heat of fusion 247.6 kJ kg-1 and thermal expansion 2.67 10-4 K-1. Tin was initially solid at the melting ...<|separator|>
  15. [15]
    Paraffin as Phase Change Material - IntechOpen
    Fatty acids exhibit high stability to deformation and phase separations for many cycles and also crystallize without supercooling. Their main disadvantages are ...4. Methods For Using... · 4.1 Encapsulation Of Ppcms · 4.2 Shape-Stable Ppcms
  16. [16]
    Thermal energy storage and thermal conductivity properties of ...
    Jun 8, 2020 · Fatty alcohols have been identified as promising organic phase change materials (PCMs) for thermal energy storage, because of their suitable ...
  17. [17]
    An overview of polyethylene glycol composite phase change materials
    Dec 20, 2024 · Polyethylene glycol (PEG) based organic phase change materials (PCMs) have great potential in the fields of solar energy, thermal management, etc.
  18. [18]
    Organic Phase Change Material - an overview | ScienceDirect Topics
    Organic solid-liquid phase change materials are mainly composed of fatty acids, fatty hydrocarbons, alcohols, etc. They have the advantages of good molding, no ...
  19. [19]
    Thermal energy storage and thermal conductivity properties of fatty ...
    Sep 21, 2020 · Like all other organic PCMs, FAs also suffer from low thermal conductivity (TC) and leakage issue, if not properly housed. Easy and cheaper ...
  20. [20]
    Evaluation of volume change in phase change materials during their ...
    This paper analyses the volume change of four organic PCM with different nature (paraffin, polymer, sugar alcohol, and aromatic hydrocarbon) in the laboratory ...
  21. [21]
    What about greener phase change materials? A review on biobased ...
    The present paper reviews the current research status of biobased PCMs with focus on various limitations and drawbacks, and on strategies adopted for improving ...
  22. [22]
    Sustainable Organic Phase Change Materials for ... - PubMed Central
    May 14, 2025 · The waste fat, composed of both saturated and unsaturated fatty acids, presented a melting temperature of 32 °C, i.e., still suitable for indoor ...
  23. [23]
    Bio-Based Phase Change Materials for Sustainable Development
    Sep 30, 2024 · This paper discusses the state of the art of bio-based phase change materials (bio-PCMs), derived from animal fats and plant oils as sustainable alternatives.
  24. [24]
    Salt hydrate phase change materials: Current state of art and the ...
    These phase change materials have better thermal performance, better flame retardance, lower manufacturing costs, and a more sustainable supply than their ...
  25. [25]
    Thermal Storage of Nitrate Salts as Phase Change Materials (PCMs)
    Nov 26, 2021 · According to Li et al., the enthalpy of fusion and the melting temperature of the sodium nitrate are 178.6 kJ/kg and 306.4 °C, respectively [4].
  26. [26]
    Optimized Preparation of a Low-Working-Temperature Gallium ...
    Aug 7, 2022 · It has a melting point of 29.8 °C and latent heat of 80.1 J·g–1; with the high density of Ga, the volume heat storage density can reach 488.2 J ...
  27. [27]
    Advancements and challenges in enhancing salt hydrate phase ...
    Salt hydrate PCMs exhibit incongruent melting during the endothermic process, leading to phase separation during multiple thermal cycles, as shown in Figure 13.
  28. [28]
    A eutectic salt high temperature phase change material
    The melting temperature and latent heat of NaClNa2CO3 eutectic salt are 635.0 °C and 311.6 J/g, and after 1000 thermal cycles, those do not change significantly ...
  29. [29]
    Synthesis of solid–solid phase change material for thermal energy ...
    Jul 19, 2025 · Polyethylene glycol (PEG10000)/ poly (glycidyl methacrylate) (PGMA) crosslinked copolymer was prepared as solid-solid phase change material and ...
  30. [30]
    Cross-linked polyurethane as solid-solid phase change material for ...
    High latent heat reaching up to 116 ± 1 J/g can be achieved with transition temperatures of 58.5 °C (for melting) and 37 °C (for crystallisation); (ii) Form- ...
  31. [31]
  32. [32]
    Thermal energy storage performance, application and challenge of ...
    Phase change material (PCM) has critical applications in thermal energy storage (TES) and conversion systems due to significant capacity to store and ...
  33. [33]
    Recent developments in thermo-physical property enhancement ...
    Aug 28, 2019 · This review addresses current research on SS-PCM with the focus on solid solid phase transition mechanisms, properties, thermo-physical property enhancement,<|separator|>
  34. [34]
    Eutectic mixtures of capric acid and lauric acid applied in building ...
    The results showed that the melting temperature and latent heat of these phase change wallboards with eutectic mixtures have not obvious variations after ...
  35. [35]
    Thermal Properties and Reliabilities of Lauric Acid-Based Binary ...
    Apr 27, 2022 · The eutectic PCMs have a melting enthalpy of 166.18, 183.07, and 189.50 J·g–1 and a melting temperature of 35.10, 37.15, and 39.29 °C for lauric ...
  36. [36]
    [PDF] Phase Change Material (PCM) as the Smart Heat-Storing Concept
    Dec 27, 2022 · The attractive identities of PCM materials are high capacity of thermal energy storage, great heat conductivity, little dilatation, shrinkage ...
  37. [37]
    [PDF] Characterization of Thermophysical Properties of Phase Change ...
    Sep 9, 2020 · The most relevant properties in the characterization of phase change materials (PCM) are the phase change enthalpy, thermal conductivity, heat ...<|control11|><|separator|>
  38. [38]
    [PDF] Recent advances in thermophysical properties enhancement ... - HAL
    Aug 22, 2023 · properties of PCM in order to promote its use as an efficient thermal energy storage medium. 979. Overall, the findings of the literature review ...
  39. [39]
    Phase change material-based thermal energy storage - ScienceDirect
    Aug 18, 2021 · Phase change materials (PCMs) having a large latent heat during solid-liquid phase transition are promising for thermal energy storage applications.Missing: Storness | Show results with:Storness
  40. [40]
    Selection of the Appropriate Phase Change Material for Two ... - MDPI
    Mar 20, 2020 · The aim of this paper is to present the methodology used to assess the suitability of potential phase change materials to be used in two innovative energy ...
  41. [41]
    Investigation on compatibility and thermal reliability of phase change ...
    Nov 22, 2020 · The PCM selection is made by targeting low temperature (<100 °C) heat storage applications. The PCMs considered are paraffin wax, sodium acetate ...Experimental Methodology · Corrosion Study · Pcm Behavior
  42. [42]
    PCM Selection for Heat Pump Integrated with Thermal Energy ...
    Jul 1, 2022 · This paper will present an analysis of the PCM material selection for heat pump integrated TES in various configurations.
  43. [43]
    Tests on Material Compatibility of Phase Change Materials and ...
    Apr 10, 2019 · The results of compatibility tests are clear. Selection of proper PCM-capsule pairs is crucial for long-term applications in buildings. The ...
  44. [44]
    The performance evaluation of shape-stabilized phase change ...
    ESI and ESE can evaluate a passive component/material from an energy standpoint. •. A kind of PCM's performance is demonstrated in a full size lightweight room.
  45. [45]
    Optimizing the design of composite phase change materials for high ...
    Oct 10, 2018 · (a) The phase diagram for the figure-of-merit ηeff,max is calculated using Eq. (6) and is a function of the thermal conductivity ratio κ = kC/k ...Ii. Material Design Space · B. Composite Energy Density · C. Composite Thermal...
  46. [46]
    Selection of phase‐change material for thermal management of ...
    Sep 14, 2020 · The present study considers 30 PCMs and 11 attributes for the selection of best PCM. The results are compared with both the methods and it is observed that RT ...
  47. [47]
    Review on the sustainability of phase-change materials used in ...
    In this study, we perform a review on the sustainability of phase-change materials considering performance, economic, environmental, and social aspects.
  48. [48]
    Joseph Black and Latent Heat - American Physical Society
    The latent heat that Black discovered greatly slows the melting of snow and ice. He gave the first account of this work on April 23, 1762 at the University of ...
  49. [49]
    The Sun Queen and the Skeptic: Building the World's First Solar ...
    Jul 14, 2020 · More promising were phase-change materials that absorb or release heat when they change from solid to liquid. She became fascinated with one ...
  50. [50]
    The 1948 Dover Sun House Used Phase Change Materials to Store ...
    Chemical engineer Maria Telkes designed a system relying on a phase-changing material, Glauber's Salt or the decahydrate of sodium sulfate.
  51. [51]
    Advancement in thermal energy storage using phase change ...
    Advancement in thermal energy storage using phase change materials. Author. Rami Mohammad Reda Saeed. Keywords and Phrases. Energy Storage Materials; Nuclear ...
  52. [52]
    [PDF] TECHNOLOGY - NASA Technical Reports Server
    In other studies,. Lockheed thoroughly investigated phase change materials (PCMs) for high- capacity energy storage systems. In still other NASA-funded studies, ...
  53. [53]
    Introductory Chapter: Phase Change Material - IntechOpen
    The discovery of phase change material (PCM) starts from the early 1900s in the work of Alan Tower Waterman of Yale University [1]. While studying thermionic ...
  54. [54]
    None
    Below is a merged summary of the history of Phase Change Materials (PCMs) based on the provided segments. To retain all information in a dense and organized manner, I will use a table in CSV format to capture the details across the different summaries, followed by a concise narrative overview that integrates the key points. The table will include all segments' contributions for each category (e.g., Early History, Military Applications, etc.), ensuring no information is lost.
  55. [55]
    PCM products and their fields of application - ScienceDirect.com
    Phase Change Materials, or briefly PCM, are a promising option for thermal energy storage, depending on the application also called heat and cold storage.
  56. [56]
    Advancement in Thermal Energy Storage Using Phase Change ...
    Advancement in Thermal Energy Storage Using Phase Change Materials. March 2018. Thesis for: DOCTOR OF PHILOSOPHY IN NUCLEAR ENGINEERING; Advisor: Joshua P.
  57. [57]
    Macrocapsules containing microencapsulated phase change ...
    The present invention relates to thermal energy storage compositions comprising macrocapsules or beads containing phase change materials (PCMs) and, more ...
  58. [58]
    Microencapsulation of higher hydrocarbon phase change materials ...
    Aug 7, 2025 · PCMs is always microencapsulated, and that has been studied by many researchers [7][8][9][10] [11] and developed by some companies like BASF.
  59. [59]
    [PDF] A History of the Energy Research and Development Administration
    The Arab oil embargo of October 16, 1973, had an immediate impact on the United States. On. November 8 President Nixon sent a message to Congress stating that ...
  60. [60]
  61. [61]
    Preparation and characterization of paraffin/expanded graphite ...
    Dec 1, 2022 · The results show that the thermal conductivity of PCM/EG increases with EG content and block density increasing. When EG content is 20%, ...
  62. [62]
    Thermally Conductive Shape-Stabilized Phase Change Materials ...
    Compared to paraffin wax, the obtained paraffin wax/expanded graphite composites have about 20 times higher thermal conductivity with less than 13% loss of ...
  63. [63]
    Form-stable paraffin/high density polyethylene composites as solid ...
    This paper deals with the preparation of paraffin/high density polyethylene (HDPE) composites as form-stable, solid–liquid phase change material (PCM) for ...
  64. [64]
    Thermal conductivity measurement of a PCM based storage system ...
    Aug 9, 2025 · The carbon fibers essentially enhance the effective thermal conductivity of fiber/paraffin composites. For the random type, the fiber length ...
  65. [65]
    Metal foams application to enhance the thermal performance of ...
    MFs can significantly improve the thermal performance of PCMs by offering a highly conductive and sturdy structure.
  66. [66]
    Mathematical model for evaluating effective thermal conductivity of ...
    The two basic equations of thermal conduction used in composites are Rule of Mixture ... composite PCM are sought to achieve high energy charging ...
  67. [67]
    Experimenta l study on the thermal stability of a paraffin mixture with ...
    Aug 10, 2025 · The experimental results presented that the degeneration of the heat of fusion was 9.1% after 10000 thermal cycles while that of the melting ...
  68. [68]
    Phase Change Materials Composite Based on Hybrid Aerogel with ...
    Feb 7, 2021 · Phase change materials (PCMs) can be thermally enhanced by reduced graphene oxide (rGO)/expanded graphite (EG) aerogel with anisotropic microstructure.
  69. [69]
    Composite phase-change materials for photo-thermal conversion ...
    Jun 1, 2024 · The OBC/GNP/PA PCM composites with 4 wt% GNP show a photo-thermal conversion efficiency of 86.21%. The thermal conductivity increases by 155% ...
  70. [70]
    Photothermal Nanomaterials: A Powerful Light-to-Heat Converter
    May 3, 2023 · We review the latest progresses on photothermal nanomaterials, with a focus on their underlying mechanisms as powerful light-to-heat converters.Missing: PCM | Show results with:PCM
  71. [71]
    Photothermal Phase Change Energy Storage Materials
    Aug 20, 2024 · To meet the demands of the global energy transition, photothermal phase change energy storage materials have emerged as an innovative solution.
  72. [72]
    MXene Ti3C2Tx for phase change composite with superior ...
    The composite has strong absorption, enhanced by LSPR, and photo-to-thermal storage efficiencies up to 94.5%, high energy storage density, and form-stability.
  73. [73]
    Sandwich-structured MXene/wood aerogel with waste heat ...
    Feb 15, 2023 · Sandwich-structured MXene/wood aerogel coupled with PCMs evaporator is developed. · Evaporation efficiency is 92.6% under 1 sun irradiation with ...
  74. [74]
    Potential Phase Change Materials in Building Wall Construction—A ...
    Microencapsulated PCMs (e.g., paraffin) can be mixed directly with gypsum and used in the form of a panel, as shown in Figure 16 from Micronal® produced by ...
  75. [75]
    Enhancing energy performance of glazing systems using solid-solid ...
    Jan 18, 2025 · This study explores the unique combination of solid-solid phase change material (SSPCM) in double-glazed window (DGW) subjected to different ...
  76. [76]
    Phase Change Materials in Residential Buildings - PubMed Central
    The results highlight that optimized PCM integration can reduce energy consumption by up to 30% and improve indoor thermal comfort. However, challenges such as ...
  77. [77]
    Implementation of phase change material for cooling load reduction
    Oct 30, 2022 · This research examines the impact of PCM integrated into a traditional wall in Egypt on peak and average cooling energy consumption.
  78. [78]
    (PDF) Experimental Study on the Dynamic Integration of Pcm in ...
    Experimental Study on the Dynamic Integration of Pcm in Building Walls for Enhancing Thermal Performance in Summer Condition. January 2024.
  79. [79]
    [PDF] Using Phase Change Materials (PCM) to Reduce Energy ...
    Using BioPCM. M91/Q23 can significantly reduce heating and cooling costs. The article concludes that integrating PCM into building components can improve indoor ...
  80. [80]
    [PDF] Application of Phase-Change Materials in Buildings: Case Study Al ...
    This paper studies the potential application of PCMs in building energy conservation materials. The analysis shows that the use of BioPCM material as an.
  81. [81]
    Integrating phase change materials in buildings for heating and ...
    This study aims to obtain and analyse the global performance of integrating PCMs in building envelops for heating and cooling demand reduction. It directly ...
  82. [82]
    [PDF] Data sheet - RT42 - Rubitherm Technologies GmbH
    Specific heat capacity. [kJ/kg·K]. Combination of latent and sensible heat in a temperatur range of °C to °C. 35. 50 main peak: 41. 42 main peak: Typical Values.
  83. [83]
    Carbon-Enhanced Hydrated Salt Phase Change Materials for ... - NIH
    Jun 24, 2024 · In many cases, semi-congruent or incongruent melting occurs, meaning that the hydrated salts are not fully dehydrated in the single step and ...
  84. [84]
    A hybrid thermal management system for high power lithium-ion ...
    In this paper, a hybrid TMS (HTMS) using phase change materials (PCM) and six flat heat pipes is proposed to maintain the temperature profile below 40 °C.
  85. [85]
    Hybrid Heat Pipe-PCM-Assisted Thermal Management for Lithium ...
    The experimental setup includes a heater section, a phase change material (PCM) reservoir, and a cooling section. The heater section simulates battery heat ...
  86. [86]
    PCM assisted heat pipe cooling system for the thermal management ...
    PCM assisted heat pipes cooling system decreases the T1 temperature to 30.73 °C. It is found that the T1 temperature decreased by 14.5% and 43.3% for the cell ...
  87. [87]
    Effective energy use and passive cooling in data centers using heat ...
    Jun 30, 2025 · Passive cooling with TEG-PCM-heat pipes recovers data center waste heat. · TEG-PCM coupling ensures stable power output, novel for data centers.
  88. [88]
    Phase Change Fabrics Control Temperature | NASA Spinoff
    Phase change materials (PCMs) control temperature swings in textiles using passive control strategies. These PCMs change state at different temperatures.
  89. [89]
    Characterization of the PCM fibers - Textile Today
    Dec 12, 2015 · Outlast technology is the leading developer of the PCM technology. The ... Thermasorb 83 (Outlast technologies, 28°c). ii. Thermasorb 95 ...Missing: point | Show results with:point
  90. [90]
    Development of composite phase change cold storage material and ...
    This research aims to explore a novel PCM that can satisfy the needs of vaccine cold chain logistics temperature zone (2–8 °C), with decyl alcohol (DA) and ...Missing: reefer trucks
  91. [91]
    Thermal Performance of Cold Panels with Phase Change Materials ...
    Applying the Phase Change Materials (PCMs) in the refrigerated truck is a reliable method of cold storage because they have high latent heat of melting and ...
  92. [92]
    Sustaining Vaccine Potency in Cold Chain Logistics - MDPI
    Vaccination cold chains depend critically on maintaining temperatures within the 2–8 °C range, with phase-change materials (PCMs) like n-tetradecane ...Missing: reefer trucks
  93. [93]
    Review on the challenges of salt phase change materials for energy ...
    Feb 1, 2024 · This review summarises new advancements in phase change material research, a comparison analysis of salts and other storage technologies, and recommendations ...
  94. [94]
    [PDF] Encapsulated phase change material for high temperature thermal ...
    Thermal analysis of high temperature phase change materials (PCM) is conducted with the consideration of a 20% void and buoyancy-driven convection in a ...Missing: ESI = ΔHf / ΔT
  95. [95]
    Active implantable drug delivery systems: engineering factors ...
    Jul 10, 2025 · The device features a refillable microreservoir and a peristaltic micropump based on phase-change materials, suitable for subcutaneous implants ...
  96. [96]
    A Temperature-Sensitive Drug Release System Based on Phase ...
    We have fabricated PCM-based drug delivery constructs in the forms of spheres and blocks to demonstrate on-demand drug release as regulated by temperature. In ...
  97. [97]
    A phase-change-material-assisted absorption thermal battery for ...
    Jul 15, 2024 · This study proposed a Phase-Change-Material (PCM)-assisted ATB to recover and store the condensation heat in the charging process.<|separator|>
  98. [98]
    ICSC 1457 - PARAFFIN WAX - INCHEM
    Melting point: 50-57°C Solubility in water: none. Flash point: 199°C c.c.. EXPOSURE & HEALTH EFFECTS. Routes of exposure. The substance can be absorbed into the ...
  99. [99]
    Flame Retardant Paraffin-Based Shape-Stabilized Phase Change ...
    May 21, 2020 · For PCM, the heat release rate (HRR) increased rapidly after ignition and the peak heat release rate (PHRR) reached 1137.0 kW/m2 in 145 s.
  100. [100]
    Fire Retardant Phase Change Materials—Recent Developments ...
    Jun 14, 2023 · The cone calorimetry test indicated that the peak heat release rate decreased from 870.9 kW/m2 to 313.1 kW/m2 after the addition of flame ...Missing: flash | Show results with:flash
  101. [101]
    Cone Calorimetry in Fire-Resistant Materials - IntechOpen
    Jun 23, 2022 · Cone Calorimetry is the most suitable test method for predicting the real-scale fire behavior of polymeric materials.2. Thermal Decomposition Of... · 4. Types Of... · 4.4 Cone Calorimetery
  102. [102]
    Cycle test stability and corrosion evaluation of phase change ...
    One of the critical problems with the PCM is its degradation with the continuous melt/freeze cycles. Because of that, the PCM loses its heat storage capability ...
  103. [103]
    Prospects and challenges of bio-based phase change materials
    Jun 15, 2024 · Some drawbacks of bio-based PCMs such as degradation due to microorganisms, thermal and chemical instabilities, and degradation due to cycling ...
  104. [104]
    Fire Retardance Methods and Materials for Phase Change ... - MDPI
    Apr 23, 2023 · This review paper covers current studies assessing the PCM response to fire or excessive temperature, methods for ensuring flame retardancy, and their impact ...2.1. Ul 94 Horizontal... · 3. Fire Retardants... · 3.4. Flame Retardant Surface...
  105. [105]
    Thermal stability of organic Phase Change Materials (PCMs) by ...
    A minimum of 10,000 thermal cycles positions the PCM favourably for applications in the built environment. In accordance with British Standard for reinforced or ...
  106. [106]
    Phase Change Materials in Metal Casting Processes: A Critical ...
    Feb 4, 2022 · Despite having advantages, inorganic PCM comes with major disadvantages, that is, toxicity, corrosivity, supercooling, and low thermal expansion ...Missing: heavy | Show results with:heavy
  107. [107]
    A form-stable phase change material based on intermolecular ...
    Thus, PEG and mannitol are obtained, with recycling rates greater than 90% for each. Additionally, there are high latent heat values for this PEG-based ...
  108. [108]
  109. [109]
    [PDF] PCM Price challenge | HubSpot
    The PCM material is prepared in-house and contains no impurities. The cost of the raw material used is $0.90−1.80/lb ($1.98−3.96/kg). The PCM products are in ...
  110. [110]
    Energy-carbon-investment payback analysis of prefabricated ...
    The carbon and energy payback periods show the same trend (3–7 years); the economic payback periods (5–30 years) are higher than the environmental periods.
  111. [111]
    Phase Change Materials Market Size, Share & Trends Report, 2034
    The global phase change materials market was estimated at USD 2.6 billion in 2024. The market is expected to grow from USD 3 billion in 2025 to USD 7.9 billion ...
  112. [112]
    Advances in the development of latent heat storage materials based ...
    The main objective of this study is to discuss lithium compounds used, proposed or analyzed for latent heat storage (LHS) and their possible applications.Lithium In Latent Tes · Lithium Salt Applications In... · Economic Costs Of Li Based...Missing: chain | Show results with:chain
  113. [113]
    REACH Regulation - Environment - European Commission
    REACH is the main EU law to protect human health and the environment from the risks that can be posed by chemicals.
  114. [114]
    Life-Cycle Assessment of phase-change materials in buildings
    Feb 15, 2022 · LCA is a well-established method for quantification of environmental impacts (resource consumption, emissions, etc.) of products during their ...Missing: CO2 | Show results with:CO2
  115. [115]
    Metal–Organic Network-Based Composite Phase Change Materials ...
    In addition, the photothermal conversion efficiency of PEG/CFK-2 was calculated to be 83.4% using formula (1). The above results suggested that PEG/CFK-2 ...2. Materials And Methods · 3. Results And Discussion · 3.4. Photothermal Conversion...
  116. [116]
    Graphene-Enhanced Composite Phase Change Materials for ...
    A paraffin/graphene composite with hyperbolic graphene aerogel (HGA) achieved thermal conductivity up to 82.4 W/mK with 30 wt% HGA. A 3D graphene skeleton ( ...
  117. [117]
    Design of Sustainable Self‐Healing Phase Change Materials by ...
    Oct 31, 2023 · The sustainable self-healing phase change materials (SHPCMs) are designed by introducing sustainable fatty alcohols as phase change ...Missing: encapsulation microcapsules
  118. [118]
    Recent advances in phase change microcapsules for oilfield ...
    Apr 7, 2025 · This review highlights the critical role of phase change microcapsules in advancing thermal management solutions for the efficient development ...
  119. [119]
    Recent advances in the properties, synthesis, and applications of ...
    Oct 8, 2025 · This review highlights advanced oriented phase change composites (OCPCMs) that achieve exceptional directional heat transfer and high energy ...
  120. [120]
    Advances in encapsulated phase change materials for integration in ...
    Jun 24, 2025 · This review also discusses key properties and proposed figure of merit for ePCMs, impact of encapsulation on thermophysical properties ...
  121. [121]
    Recent Advances in Phase Change Energy Storage Materials ...
    Jan 22, 2025 · This paper offers a thorough examination of the latest developments in PCES materials (PCESMs) and their wide-ranging applications in several industries.
  122. [122]
  123. [123]
    Clinic-on-a-Needle Array toward Future Minimally Invasive ...
    Jun 22, 2021 · coated with phase change materials. This enables multistage, spatially patterned, and precisely controlled drug release in response to the ...
  124. [124]
  125. [125]
    Laboratory Experiments on Passive Thermal Control of Space ...
    Here, we investigate the performance of phase-change materials (PCMs) in the passive thermal control of space habitats.
  126. [126]
    Phase change materials for enhanced photovoltaic panels ...
    This research article shows the potential of PCM-based cooling solutions in advancing renewable energy technologies and covers a comprehensive review
  127. [127]
    Phase Change Materials in Residential Buildings - MDPI
    Phase change materials (PCMs) have emerged as promising solutions for improving thermal management in residential buildings by enhancing thermal storage ...
  128. [128]
    The contribution of artificial intelligence to phase change materials ...
    This comprehensive review delves into AI applications within the domain of PCM for TES systems, mainly including prediction and optimization.<|control11|><|separator|>
  129. [129]
    A review on advanced phase change material-based cooling for ...
    PCMs offer a passive and sustainable cooling solution, enhancing system reliability and energy efficiency. To maximize thermal energy management in electronic ...
  130. [130]
    Recent advances in the properties, synthesis, and applications of ...
    Phase Change Materials. Article. Recent advances in the properties, synthesis, and applications of oriented composite phase change materials. September 2025 ...