Fact-checked by Grok 2 weeks ago

Enthalpy of fusion

The of fusion, denoted as ΔH_fus, is the change in associated with the of a substance from solid to at its and constant , without any accompanying change, representing the energy required to overcome intermolecular forces in the solid lattice. This process absorbs , increasing the potential energy of the molecules while kinetic energy remains constant. It is typically reported as a quantity in units of kilojoules per (/), though specific enthalpies per unit mass (/kg or J/g) are also common for practical applications. A well-known example is , where the molar enthalpy of fusion is 6.00678 kJ/mol at 0°C and 101.325 kPa, equivalent to approximately 333.55 J/g for melting into liquid . This value reflects the relatively strong bonding in , which requires significant to disrupt. In general, the magnitude of ΔH_fus correlates with the strength of intermolecular forces; substances with stronger bonds, such as metals or ionic solids, exhibit higher values compared to molecular solids with weaker van der Waals interactions.

Core Concepts

Definition

The enthalpy of fusion, denoted as \Delta H_\text{fus}, is the change in enthalpy accompanying the phase transition of one mole of a substance from the solid to the liquid state at constant pressure and its melting temperature T_m. This transition is an endothermic process in which heat is absorbed to overcome intermolecular forces, disrupting the ordered solid structure without altering the temperature of the substance during the melting phase. The enthalpy of fusion is synonymous with the of fusion for the process at constant , where the absorbed equals \Delta H_\text{fus}, as \Delta H = q_p. Mathematically, it is expressed as \Delta H_\text{fus} = H_\text{[liquid](/page/Liquid)} - H_\text{[solid](/page/Solid)} evaluated at T_m. The standard units are kilojoules per mole (kJ/) for molar quantities or joules per gram (J/g) for specific values, with the SI unit being joules per (J/).

Thermodynamic Interpretation

The enthalpy of fusion, denoted as \Delta H_\text{fus}, represents the absorbed during the from to at constant and is fundamentally linked to of . According to the definition of as H = U + PV, where U is the , P is , and V is , the change in for the process is given by \Delta H_\text{fus} = \Delta U_\text{fus} + P \Delta V_\text{fus}. Here, \Delta U_\text{fus} accounts for the change in molecular interactions and vibrational freedom as the ordered structure breaks into the more disordered , while P \Delta V_\text{fus} captures the work associated with the typically small expansion upon , since liquids generally occupy slightly more space than for most substances. From the perspective of the second law of thermodynamics, the enthalpy of fusion connects to the Gibbs free energy change at the equilibrium melting temperature T_m, where \Delta G_\text{fus} = 0. This condition implies \Delta G_\text{fus} = \Delta H_\text{fus} - T_m \Delta S_\text{fus} = 0, yielding the key relation \Delta S_\text{fus} = \Delta H_\text{fus} / T_m, which quantifies the entropy increase due to the greater disorder in the liquid phase compared to the solid. Unlike the entropy of vaporization, which follows Trouton's rule with a roughly constant value of about 85–88 J/mol·K for many non-associated liquids, the entropy of fusion varies more widely (typically 10–60 J/mol·K depending on the substance type, such as metals or molecular solids) but adheres to the same thermodynamic equality at equilibrium. In phase diagrams, the enthalpy of fusion plays a critical role in determining the slope of the solid-liquid equilibrium line through the Clapeyron equation, derived from the equality of chemical potentials across phases: \frac{dT}{dP} = \frac{T_m \Delta V_\text{fus}}{\Delta H_\text{fus}}. This equation illustrates how influences the ; for most where \Delta V_\text{fus} > [0](/page/0), increasing raises T_m, as the denominator \Delta H_\text{fus} (always positive for endothermic melting) moderates the effect alongside the volume change. Exceptions occur for like , where \Delta V_\text{fus} < [0](/page/0), leading to a decrease in with . For reversible fusion processes at constant pressure, the heat absorbed q_\text{rev} equals the enthalpy change \Delta H_\text{fus}, which also matches T_m \Delta S_\text{fus} due to the reversible nature of the phase transition at equilibrium. This equivalence underscores the process's isothermal character, where the system absorbs latent heat without temperature variation, balancing the entropy production to zero for the universe.

Measurement and Data

Experimental Determination

The experimental determination of the enthalpy of fusion, denoted as \Delta H_\text{fus}, traces its origins to the late 18th century, when Antoine Lavoisier and Pierre-Simon Laplace developed an ice calorimeter in 1782 to measure latent heats relative to the heat required to raise water from 0°C to 60°C. In their apparatus, a sample was enclosed within concentric ice-filled containers; heat absorbed or released by the sample melted a quantifiable amount of ice (approximately 489.5 g per "pound" equivalent), allowing relative determinations of specific and latent heats with an estimated accuracy of about 1.7%, though modern recalibrations show deviations of 10-15% for some values. Classical methods for measuring \Delta H_\text{fus} relied on calorimetry via the method of mixtures, where a known mass of solid at or below its melting point is added to a calorimeter containing a liquid (often water) initially above the melting temperature, and the equilibrium temperature change is used to calculate the latent heat absorbed during melting. This approach quantifies heat input by monitoring temperature equilibration, assuming no heat loss and complete phase change, and has been applied to substances like ice since the 19th century. Post-1900, these techniques evolved through the establishment of absolute calorimetric standards by national metrology institutes, incorporating electrical calibration via the for improved traceability to SI units and greater precision in heat flux measurements. Differential scanning calorimetry (DSC) serves as the modern standard for determining \Delta H_\text{fus}, operating by heating a sample and reference at a controlled rate while measuring differential heat flow to maintain identical temperatures. The melting process produces an endothermic peak on the heat flow versus temperature plot; \Delta H_\text{fus} is obtained by integrating the peak area between the initial onset (where the curve deviates from the baseline) and final temperature (where it returns to baseline), often normalized to sample mass and calibrated against standards like indium (\Delta H_\text{fus} = 28.47 J/g at 156.4°C). This method is applicable to thermally stable materials over -120°C to 600°C, providing rapid results for quality control and research, though outcomes can vary with sample form and heating rate. For precise low-temperature measurements, adiabatic calorimetry isolates the sample from external heat exchange, allowing accurate tracking of heat capacity and phase transitions by incrementally adding electrical energy while maintaining near-zero temperature gradients. This technique has been used to determine \Delta H_\text{fus} for metals like gallium, yielding values such as 5.59 kJ/mol at 29.78°C for high-purity samples. Key challenges in these measurements include supercooling, where the liquid phase persists below the melting point without crystallizing, leading to underestimation of \Delta H_\text{fus} if the end of the latent heat period is misidentified. Impurities depress the melting point and reduce the observed \Delta H_\text{fus} due to eutectic formation or incomplete phase purity, with effects quantifiable via DSC purity analysis comparing experimental onsets to theoretical 100% purity values. Additionally, incomplete melting—arising from kinetic barriers or insufficient heat supply—results in lower calculated enthalpies, as unmelted fractions do not contribute to the full endothermic response.

Tabulated Values and Examples

The enthalpy of fusion, denoted as ΔH_fus, varies significantly across substances depending on their bonding type, molecular structure, and atomic mass. Representative values for elements illustrate differences between noble gases, alkali metals, and transition metals, with noble gases exhibiting particularly low values due to weak interatomic forces. For instance, helium has an anomalously small ΔH_fus of 0.02 kJ/mol, reflecting its quantum mechanical behavior near absolute zero, while metals like sodium show lower values compared to heavier transition metals like iron.
ElementΔH_fus (kJ/mol)Melting Point (°C)
Helium (He)0.02-272 (under pressure)
Sodium (Na)2.6097.8
Iron (Fe)13.81538
Data from CRC Handbook of Chemistry and Physics (92nd ed.) and standard references. For compounds, molecular solids like water and naphthalene have moderate ΔH_fus influenced by hydrogen bonding or van der Waals forces, whereas ionic compounds like sodium chloride require higher energy to disrupt lattice structures. Organic compounds such as glucose exhibit higher values due to extensive hydrogen bonding in their crystal forms. These values are typically reported at standard pressure and the substance's melting point, with units in kJ/mol for molar quantities.
CompoundΔH_fus (kJ/mol)Melting Point (°C)
Water (H₂O)6.010.00
Naphthalene (C₁₀H₈)18.880.2
α-D-Glucose (C₆H₁₂O₆)31.4141
Sodium chloride (NaCl)28.2801
Data from CRC Handbook of Chemistry and Physics (92nd ed.) and NIST Chemistry WebBook. Trends in ΔH_fus reveal that ionic solids generally have higher values (e.g., NaCl at 28.2 kJ/mol) than molecular solids (e.g., naphthalene at 18.8 kJ/mol), as breaking electrostatic bonds requires more energy than overcoming weaker intermolecular forces. Metals often fall in between, with lighter alkali metals like sodium showing lower ΔH_fus (2.60 kJ/mol) due to metallic bonding delocalization, while heavier metals like iron require more (13.8 kJ/mol). Anomalies include helium's near-zero value (0.02 kJ/mol), associated with its superfluid transition rather than classical melting, highlighting quantum effects at low temperatures. Values can vary with sample purity, crystal form, and isotopic composition; for example, heavy water (D₂O) has a slightly higher ΔH_fus (6.4 kJ/mol) than H₂O due to stronger bonding. In everyday phenomena, the enthalpy of fusion of ice (334 J/g or 6.01 kJ/mol) governs processes like melting snow, where 334 J of heat is needed per gram at 0°C, influencing climate and refrigeration cycles. Industrially, values for alloys like iron-based steels (around 13.8 kJ/mol for pure Fe, adjusted for alloys) are critical in metallurgy for energy calculations in smelting and casting, ensuring efficient phase transitions without excessive thermal input.

Applications and Models

Solubility Prediction

The enthalpy of fusion plays a central role in thermodynamic models for predicting the solubility of solid solutes in liquid solvents, particularly through equations that relate the equilibrium between the solid and dissolved phases. For ideal solutions, where the solute behaves as if it were dissolving in itself (i.e., no heat of mixing and activity coefficient of unity), the mole fraction solubility x at temperature T is given by the : \ln x = -\frac{\Delta H_{\fus}}{R} \left( \frac{1}{T_{\mathrm{m}}} - \frac{1}{T} \right) Here, \Delta H_{\fus} is the molar enthalpy of fusion, T_{\mathrm{m}} is the normal melting temperature of the solute, R is the gas constant, and the equation assumes negligible heat capacity difference (\Delta C_p = 0) between the solid and liquid states. This simplified form arises from equating the chemical potentials of the pure solid and the hypothetical supercooled liquid solute in ideal solution. The Schröder–van Laar equation originated as an adaptation of the for the temperature dependence of solubility in the late 19th and early 20th centuries, initially proposed by Schröder in 1891 for ideal cases and extended by van Laar around 1910 to broader phase equilibria. These developments built on van't Hoff's foundational work on osmotic pressure and equilibrium constants, applying it to solid-liquid dissolution to enable predictive calculations from melting properties alone. For non-ideal solutions, where interactions between solute and solvent lead to deviations from ideality, the equation is extended by incorporating the activity coefficient \gamma of the solute, yielding \ln (x \gamma) = -\frac{\Delta H_{\fus}}{R} \left( \frac{1}{T_{\mathrm{m}}} - \frac{1}{T} \right). Activity coefficients are often estimated using group contribution methods like (UNIversal Functional Activity Coefficient), which decomposes molecules into functional groups to predict non-ideal behavior without extensive experimental data. Additionally, if \Delta C_p \neq 0, a more complete form integrates the heat capacity term: \ln (x \gamma) = -\frac{\Delta H_{\fus}}{R} \left( \frac{1}{T_{\mathrm{m}}} - \frac{1}{T} \right) - \frac{\Delta C_p}{R} \left( \ln \frac{T}{T_{\mathrm{m}}} - \frac{T}{T_{\mathrm{m}}} + 1 \right), accounting for temperature-dependent enthalpies. In pharmaceutical applications, these models are widely used to predict drug solubility in organic solvents or formulations, aiding in solvent selection and process design; for instance, the ideal solubility of a compound like ibuprofen can be estimated from its \Delta H_{\fus} and T_{\mathrm{m}} to guide dissolution studies. Similarly, for inorganic salts in aqueous or organic media, such as sodium chloride in ethanol-water mixtures, the approach forecasts solubility curves to optimize extraction or purification processes. Key limitations include the assumption of a single stable solid form, which fails for polymorphic drugs where different crystal structures have varying \Delta H_{\fus} and T_{\mathrm{m}}, leading to inaccurate predictions for metastable forms (e.g., up to 50% solubility overestimation if fusion properties are mismatched). Errors also arise in systems with significant \Delta C_p, particularly near the melting point, where neglecting it can introduce deviations of 10–20% in predicted solubilities.

Materials and Phase Behavior

In materials science and engineering, the enthalpy of fusion (ΔH_fus) plays a critical role in the casting and solidification processes by governing the latent heat released during the liquid-to-solid phase transition, which must be accounted for in the overall energy balance to control cooling rates and microstructure formation. During freezing, this latent heat release slows the temperature drop, influencing the solidification front's progression and preventing defects like shrinkage porosity in castings. For instance, in steel production via continuous casting, the ΔH_fus of steel, approximately 250 kJ/kg, contributes significantly to the heat extraction requirements in the mold and secondary cooling zones, ensuring uniform solidification and minimizing cracking. In binary alloy systems, ΔH_fus directly impacts the construction and interpretation of phase diagrams, particularly through its effect on liquidus and solidus curves via thermodynamic models like the Schröder-van Laar equation, which relates melting point depression to fusion enthalpies and thus shifts eutectic and peritectic points. A higher ΔH_fus for one component steepens the liquidus slope, potentially widening the eutectic composition range where simultaneous solidification of two phases occurs at a constant temperature, optimizing alloy designs for lower melting points and improved castability. Similarly, in peritectic reactions—where a solid phase reacts with liquid to form another solid—disparities in ΔH_fus values between phases can alter the reaction isotherm's position and the resulting microstructure, such as in Cu-Zn brass alloys, affecting phase stability and mechanical properties. For cryogenic applications, materials with low ΔH_fus, such as nitrogen (approximately 25.5 kJ/kg), enable efficient phase change cooling by requiring minimal energy input for melting solid , which is used in specialized low-temperature systems like cryopumps or insulated storage to maintain sub-77 K environments with controlled heat absorption. This low latent heat facilitates rapid transitions without excessive thermal gradients, supporting applications in superconductivity testing and space simulation chambers. Phase change materials (PCMs) leveraging high ΔH_fus inorganic salts, such as with around 160 kJ/kg, are integral to thermal energy storage systems for regulating temperature in buildings and industrial processes by absorbing and releasing large amounts of heat during melting and solidification cycles. These salts' elevated fusion enthalpies provide high volumetric storage density, enabling passive thermal management in solar power plants or electronics cooling, where consistent energy buffering is essential. In modern additive manufacturing techniques like selective laser melting (SLM), ΔH_fus influences the melt pool dynamics by dictating the energy needed for powder fusion, affecting pool depth, width, and solidification speed to mitigate defects such as balling or lack-of-fusion porosity. Simulations incorporating latent heat release show that higher ΔH_fus values elongate the melt pool under laser irradiation, optimizing layer adhesion and part density in metals like stainless steel, where precise control enhances mechanical integrity.

Advanced Considerations

Pressure and Temperature Effects

The pressure dependence of the enthalpy of fusion, \Delta H_\text{fus}, arises from thermodynamic relations derived from the Clapeyron equation, which governs phase equilibria. Along the melting curve, the approximate change with pressure is given by \frac{d(\Delta H_\text{fus})}{dP} \approx T \left( \frac{d \Delta V}{dP} \right) + \Delta V, where \Delta V is the volume change upon fusion and T is the temperature. For most solids, \Delta V > 0 (liquid volume exceeds solid volume), and the term \frac{d \Delta V}{dP} is small and negative, leading to a modest increase in \Delta H_\text{fus} with pressure, typically on the order of a few J/ per . This effect is often negligible under ambient conditions but becomes relevant in high-pressure environments, such as deep geological formations. Water exhibits an anomalous behavior due to its negative \Delta V (ice is less dense than liquid water), resulting in a more pronounced pressure sensitivity. The Clapeyron equation predicts a decrease in melting temperature with increasing pressure, and consequently, \Delta H_\text{fus} for ice decreases under compression, altering energy requirements for phase changes. For instance, at pressures around 200 MPa (relevant to extreme high-pressure simulations or deep mantle conditions), the melting point drops by approximately 14 °C, influencing \Delta H_\text{fus} by up to 5% from its standard value of 6.01 kJ/mol. This anomaly has implications in climate science, where pressure-induced melting at the base of ice sheets (typically at 10–40 MPa, causing a 1–3 °C drop) facilitates basal sliding and accelerates ice flow, contributing to sea-level rise under warming conditions. Measurements confirm that such effects enhance glacier dynamics, with \Delta H_\text{fus} variations amplifying meltwater production in pressurized subglacial environments. At elevated pressures, direct measurements using diamond anvil cells (DACs) reveal significant increases in \Delta H_\text{fus} for metals. For example, in iron and other transition metals, measurements up to several GPa show \Delta H_\text{fus} rising notably from ambient values, attributed to enhanced intermolecular interactions and reduced atomic mobility under compression. These DAC techniques detect the signal during /freezing cycles, providing data crucial for modeling planetary interiors, where high-pressure controls core dynamics. The dependence of \Delta H_\text{fus} away from the T_m is described by Kirchhoff's law: \Delta H_\text{fus}(T) = \Delta H_\text{fus}(T_m) + \int_{T_m}^T \Delta C_p \, dT, where \Delta C_p = C_{p,\text{liquid}} - C_{p,\text{solid}} is the difference. For many substances, \Delta C_p > 0 (s have higher than s), so \Delta H_\text{fus} increases with above T_m and decreases below it. This integral accounts for the hypothetical extension of fusion enthalpy to superheated or supercooled states, though practical measurements are by . Representative calculations for compounds show linear approximations yielding errors under 5% when \Delta C_p is assumed . In supercooled liquids and supercritical states, determining \Delta H_\text{fus} poses significant challenges due to metastability and the absence of distinct phase boundaries. For supercooled water below 0°C, freezing releases latent heat close to standard values, but extrapolating \Delta H_\text{fus} via Kirchhoff's law requires precise \Delta C_p data, which is complicated by rapid nucleation and structural anomalies. Measurements often rely on calorimetric detection during induced freezing. In supercritical fluids, where critical points eliminate solid-liquid distinctions, \Delta H_\text{fus} concepts are hypothetical, with extensions via equations of state showing gradual enthalpy gradients rather than sharp transitions; experimental validation is hindered by the lack of observable fusion events.

Theoretical Estimations

Theoretical estimations of the enthalpy of fusion provide valuable tools for predicting this property for compounds where experimental data is unavailable, particularly novel organics and materials. These methods rely on molecular structure, thermodynamic approximations, and computational simulations to bridge data gaps without requiring direct calorimetric measurements. Group contribution methods estimate the enthalpy of fusion through additivity rules that decompose the molecule into structural fragments, each contributing a predefined increment to the total value. The approach developed by Chickos and coworkers for organic compounds incorporates corrections for ring strain, functional groups such as hydroxyl or carbonyl, and positional effects to refine predictions, achieving mean absolute errors around 10-15% for diverse hydrocarbons and derivatives. This method is particularly effective for non-polar and moderately polar organics, as it leverages extensive compilations of experimental data to parameterize group values. Quantum mechanical approaches, such as (DFT), compute the enthalpy of fusion as the energy difference between optimized crystal and liquid phase structures, often augmented by calculations to capture vibrational contributions in the solid. For instance, simulations using like PBE0 have predicted fusion enthalpies for small molecular crystals with accuracies within 5-10 kJ/mol, though liquid-phase modeling remains computationally intensive due to the need for to represent disorder. These methods excel for inorganic and metallic systems where empirical rules falter, providing insights into lattice effects and phase stability. Empirical correlations offer simple approximations linking the enthalpy of fusion to observable properties like melting temperature and molecular weight. For non-associated liquids, a widely used relation is \Delta H_\text{fus} \approx 56 \, T_m J/, where T_m is the in , reflecting an average of fusion near 56 J/· derived from extensive datasets. This correlation, akin to those explored by Yalkowsky for , performs best for flexible hydrocarbons but requires adjustments for hydrogen-bonded or rigid structures. Recent advancements in have enhanced predictive accuracy, with models trained on databases like NIST's thermochemical compilations outperforming traditional methods for polymers and complex organics. Post-2020 models, such as graph neural networks applied to protic salts, achieve root-mean-square errors below 5 kJ/mol by incorporating structural descriptors and melting points as inputs, enabling rapid screening for phase-change materials. These AI-driven approaches reflect 2025-era progress in integrating large datasets with molecular fingerprints for extrapolation to untested compounds. Despite these advances, theoretical estimations exhibit limitations, particularly poor performance for metals, ionic compounds, and highly associated systems where intermolecular forces defy simple additivity. Validation against experimental data remains essential to quantify uncertainties, as overestimations can exceed 20% in challenging cases. These methods complement the thermodynamic \Delta H_\text{fus} = T_m \Delta S_\text{fus}, emphasizing the need for reliable estimates in predictive frameworks.

References

  1. [1]
    10.10: Enthalpy of Fusion and Enthalpy of Vaporization
    Jun 25, 2023 · The heat which a solid absorbs when it melts is called the enthalpy of fusion or heat of fusion and is usually quoted on a molar basis.
  2. [2]
    Heat of Fusion - Chemistry LibreTexts
    Jan 29, 2023 · Applications. The heat of fusion process can be seen in countless applications and evidenced in the creation of many common household items.Introduction · Example 1 · Sublimation · Applications
  3. [3]
    Heat of Fusion - an overview | ScienceDirect Topics
    Heat of fusion is the energy needed to change a substance from solid to liquid at its melting point, with no temperature change.
  4. [4]
    [PDF] The latent heat of melting (fusion) of ice
    The experiment measures the latent heat of melting of ice by melting it in warm water, measuring temperature changes, and using the equation Mi.L + Mi.cw (t2 - ...
  5. [5]
    F2625 Standard Test Method for Measurement of Enthalpy of Fusion ...
    Apr 17, 2024 · Differential scanning calorimetry is an effective means of accurately measuring both heat of fusion and melting temperature. 5.2 This test ...
  6. [6]
    5.3: Enthalpy - Chemistry LibreTexts
    Jul 7, 2023 · At constant pressure, the change in the enthalpy of a system is equal to the heat flow: Δ ⁢ H = q p . Example ...Missing: q_p | Show results with:q_p
  7. [7]
    Enthalpy of Fusion - (Intro to Chemistry) - Fiveable
    Enthalpy of fusion, also known as latent heat of fusion, is the amount of energy (heat) required to transform a substance from a solid state to a liquid ...
  8. [8]
    Thermodynamics Equations
    Thermodynamics - Equations · entropy = ΔSsys + ΔS · entropy (S) S = k ln Ω Boltzmann formula · entropy change via reversible heat flow = qrev / T (constant T).
  9. [9]
    Trouton's rule - Chemistry LibreTexts
    Jan 29, 2023 · Trouton's rule can be used to estimate the enthalpy of vaporization of liquids whose boiling points are known. Experimental values vary rather ...
  10. [10]
    A brief history of the thermal properties metrology - ResearchGate
    May 12, 2025 · ... Lavoisier who developed in 1782 an ice calorimeter for the relative measurements of specific and latent heat. It was not until 1807 that the ...
  11. [11]
    Heat of Fusion - an overview | ScienceDirect Topics
    Heat of fusion is defined as the quantity of heat necessary to change 1 gram of a solid to a liquid without a temperature change, specifically for water, ...
  12. [12]
    E793 Standard Test Method for Enthalpies of Fusion and ... - ASTM
    Sep 12, 2024 · This test method describes the determination of the enthalpy (heat) of fusion (melting) and crystallization by differential scanning calorimetry (DSC).<|separator|>
  13. [13]
    Analysis of differential scanning calorimetry (DSC): determining the ...
    Sep 26, 2023 · This method provides accurate determination of thermal properties such as equilibrium melting temperature, latent heat, change of heat capacity.
  14. [14]
    Construction of an adiabatic calorimeter at low temperatures and ...
    A low-temperature adiabatic calorimeter was constructed for measuring heat capacities accurately and following enthalpy relaxations stably for a long time.
  15. [15]
    The Enthalpy of Fusion of Gallium - ResearchGate
    Aug 10, 2025 · An adiabatic calorimeter was used to measure the enthalpy of fusion of a very pure sample of gallium. The new value of the enthalpy of ...
  16. [16]
    Accuracy improvement of T-history method for measuring heat of ...
    However, some improper assumptions in the original method, such as using a degree of supercooling as the end of latent heat period and neglecting sensible heat ...Missing: challenges enthalpy
  17. [17]
    Purity and heat of fusion data for environmental standards as ...
    Values for the heat of fusion, experimental temperature onset, theoretical temperature onset for 100% pure compound, and percent purity are presented. DSC was ...Missing: effects | Show results with:effects
  18. [18]
    (PDF) Reduction of Enthalpy of Fusion and Anomalies during Phase ...
    Aug 7, 2025 · Large reduction of both freezing DeltaHf and melting DeltaHm enthalpies with decreasing the adsorbed weight has been observed. A model for the ...<|separator|>
  19. [19]
    Technical data for the element Helium in the Periodic Table
    Critical Pressure, 0.227 MPa (2.24 Atm) ; Critical Temperature, 5.19 K ; Heat of Fusion, 0.02 kJ/mol ; Heat of Vaporization, 0.083 kJ/mol ; Specific Heat, 5193.1 J ...Missing: enthalpy | Show results with:enthalpy
  20. [20]
    Enthalpy of Fusion | 117 | CRC Handbook of Chemistry and Physics
    ... Water H2O 0.00 6.01 Xenon Xe –111.75 2.27 Xenon difluoride F2Xe 129.03 16.8 Xenon tetrafluoride F4Xe 117.1 16.3 Xenon hexafluoride F6Xe 49.48 5.74 Ytterbium ...
  21. [21]
    α-D-Glucose - the NIST WebBook
    Enthalpy of fusion. ΔfusH (kJ/mol), Temperature (K), Reference, Comment. 31.420, 414. Parks and Thomas, 1934, DH. 31.42, 414. Domalski and Hearing, 1996, AC.
  22. [22]
    Water - the NIST WebBook
    Formula: H2O · Molecular weight: 18.0153 · IUPAC Standard InChI: InChI=1S/H2O/h1H2. Copy. InChI version 1.06 · IUPAC Standard InChIKey: XLYOFNOQVPJJNP-UHFFFAOYSA-NPhase change data · References
  23. [23]
    Prediction of Solubility in Nonideal Multicomponent Systems Using ...
    In this study the UNIFAC method was used to predict solubility in highly nonideal multicomponent systems in which only the solute enthalpy of fusion and melting ...
  24. [24]
    Theoretical and Practical Approaches for Prediction of Drug ...
    Aug 24, 2006 · In addition, if the enthalpy of mixing is positive, then the solubility will be reduced. Therefore, high melting point/high enthalpy of fusion ...
  25. [25]
    Estimating the Solubility of Active Pharmaceutical Ingredients Using ...
    Aug 16, 2025 · Predicting solubility using atomistic first-principle methods, that compare the chemical potential of the solid and the solvated phase, has ...
  26. [26]
    Properties: An Introduction to Grey Iron - AZoM
    An Introduction to Grey Iron ; Latent Heat of Fusion, 270, 275, kJ/kg, 116.079 ; Maximum Service Temperature, 623, 723, K · 661.73 ...Missing: enthalpy | Show results with:enthalpy
  27. [27]
    Mathematical and physical modelling of steel flow and solidification ...
    Another important issue in solidification modelling is the release of latent heat and the evaluation of solid, or liquid, fractions.
  28. [28]
    [PDF] From simple binary to complex multicomponent eutectic alloys
    Feb 27, 2021 · ... eutectic composition and eutectic temperature, TE, can be predicted using the following Schroeder-van Laar equation [57–59]: LnXi = ΔHi fusion.<|control11|><|separator|>
  29. [29]
    (PDF) Thermodynamic calculation of n-component eutectic mixtures
    Aug 6, 2025 · This paper presents a simple numerical method to calculate the eutectic mixture composition and melting temperature.
  30. [30]
    Heat of Fusion and Vaporization - gchem
    Molar ΔH (kJ/mol) ; methanol, 3.17, 35.2 ; nitrogen, 0.715, 5.60 ; sodium, 2.60, 97.42 ; water, 6.02, 40.7.
  31. [31]
    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.
  32. [32]
    [PDF] High Temperature Phase Change Materials for Thermal Energy ...
    Latent heat thermal energy storage (TES) systems using phase change materials (PCM) are useful because of their ability to charge and discharge a large amount ...
  33. [33]
    Effects of latent heat in additive manufacturing - ScienceDirect.com
    Jun 5, 2023 · We analyzed the effects of latent heat in conditions related to the additive manufacturing technology by solving the free boundary problem of a moving heat ...Abstract · Background · References (29)
  34. [34]
    Influence of Latent Heat of Fusion on the Melt Pool Shape and Size ...
    Nov 24, 2022 · The latent heat of fusion in the heat transfer process leads to melt pool elongation, as well as to a slight decrease in its width and depth.
  35. [35]
    On the pressure dependence of the heat of fusion and melting ...
    From these results, the pressure dependence of the enthalpy of fusion can be calculated via the Clausius-Clapeyron equation.
  36. [36]
    [PDF] PHYSICAL TRANSFORMATIONS OF PURE SUBSTANCES Chapter ...
    The normal melting point (under 100 kPa pressure) is 0 oC and the Enthalpy of Fusion is 6.01 kJ/mol. ... Pressure dependence of the Melting Point in Water ... The ...
  37. [37]
    Melting of ice under pressure - PNAS
    Sep 30, 2008 · We find that for pressures between 10 and 40 GPa, ice melts as a molecular solid. For pressures above ≈45 Gpa, there is a sharp increase in the slope of the ...
  38. [38]
    Growth rate of crystalline ice and the diffusivity of supercooled water ...
    We have determined the crystalline-ice growth rate and liquid-water diffusivity for temperatures between 180 and 262 K in ultrahigh-vacuum conditions.
  39. [39]
    [PDF] Estimating Solid-Liquid Phase Change Enthalpies and Entropies
    A group additivity method based on molecular structure is described that can be used ... estimate the fusion enthalpy of a compound. In addition, compounds whose ...
  40. [40]
    [PDF] Thermochimica Acta - UMSL
    This compendium summarizes the fusion enthalpies of approximately 1000 new measurements. A group additivity method developed to estimate the total phase ...
  41. [41]
    Estimating entropies and enthalpies of fusion of organic compounds
    Yalkowsky, Doaa Alantary. ... An Examination of the Thermodynamics of Fusion, Vaporization, and Sublimation of Ibuprofen and Naproxen by Correlation Gas ...Missing: empirical | Show results with:empirical
  42. [42]
    Machine learning-based predictions of melting temperature and ...
    In this study, linear and non-linear machine learning models are used to predict the melting temperature (Tm) and enthalpy of fusion (ΔHf) of 182 possible ...