Fact-checked by Grok 2 weeks ago

Melting-point depression

Melting-point depression is the reduction in the melting temperature of a substance, typically a , due to factors such as the presence of impurities, alloying, or reduction in (e.g., in ). This colligative-like effect in impure systems disrupts the orderly arrangement of the , allowing the solid to transition to the at a reduced compared to the pure substance. In addition to the lowered , impure samples typically exhibit a broader melting range, often spanning several degrees , whereas pure compounds melt sharply over a narrow of 0.5–1 °C. The underlying cause of melting-point depression lies in the of transitions. For a pure , the is the where the change (ΔG°) for the -to-liquid transition is zero, determined by the balance of (ΔH°) and (ΔS°) changes via the relation T_m = ΔH° / ΔS°. , even in small amounts, increase the of the liquid more significantly than the because the liquid's disordered accommodates foreign molecules more readily, while the 's rigid is perturbed, leading to a larger overall ΔS° and thus a lower T_m. This effect is particularly pronounced in systems approaching the eutectic point, where a specific impurity results in the minimum possible melting for the mixture. In practical applications, melting-point depression serves as a critical diagnostic tool in laboratories to evaluate the purity of synthesized compounds. A sample that melts below the literature value or over an extended indicates contamination, often from solvents, side products, or incomplete reactions, prompting further purification steps such as recrystallization. This principle extends beyond laboratories to , where controlled impurities, alloying, or nanoscale effects can intentionally depress melting points for applications in alloys, pharmaceuticals, and , though the magnitude of depression scales with impurity concentration, solubility, or .

Fundamentals

Definition and Scope

Melting point depression refers to the phenomenon where the melting temperature of a pure substance decreases upon the introduction of impurities or when the material is confined to nanoscale dimensions. Impurities disrupt the crystal lattice, lowering the temperature required for the solid-to-liquid transition. This effect is observed in various systems, including mixtures such as binary alloys where specific compositions exhibit reduced melting points, and in where size-dependent effects contribute to depression.

Thermodynamic Principles

The between the and phases of a pure substance occurs at the where the change for is zero, given by ΔG = ΔH - TΔS = 0, implying T_m = ΔH / ΔS, with ΔH as the and ΔS as the change upon melting. In the presence of impurities, this balance shifts because impurities typically incorporate more readily into the phase than the lattice, stabilizing the relative to the and lowering the at which their chemical potentials are equal. This arises from the impurity's contribution to the system's , favoring the phase at temperatures below the pure . The entropy of mixing plays a central role in this process, as impurities in the liquid increase the system's configurational disorder, quantified by ΔS_mix = -R (x_1 ln x_1 + x_2 ln x_2) for a binary ideal mixture, where x_1 and x_2 are the mole fractions of the host and impurity. This positive entropy contribution lowers the free energy of the liquid phase relative to the solid, favoring melting at reduced temperatures. In mixtures, the Gibbs phase rule, F = C - P + 2 (with C components and P phases), indicates two degrees of freedom for a binary system with two phases, allowing temperature and composition to vary while maintaining equilibrium, independent of pressure for most cases. Adaptations of the Clapeyron equation, dT/dP = ΔV / ΔS, further show that melting point depression in mixtures is typically pressure-independent due to small volume changes, focusing the effect on compositional entropy rather than pressure variations.

Historical Development

Early Observations

One of the earliest documented observations of , a phenomenon closely related to through , occurred in the late . In 1788, British physician and chemist Charles Blagden conducted experiments on the effect of dissolved salts, such as , on the freezing point of . He found that the addition of lowered the freezing point in an inverse proportion to the amount of relative to the in the solution, with solutions containing 1 part to 10 parts freezing at approximately -6°C rather than 0°C. These findings, published in the Philosophical Transactions of the Royal Society, laid foundational empirical groundwork for understanding solute-induced phase changes in solutions. In the , metallurgists and chemists began noting similar effects in solid systems, particularly that impurities in metals could lower their s, facilitating formation and . Chemists and metallurgists during the early 1800s observed how admixtures altered the physical properties of metals, including reduced temperatures in impure samples compared to pure elements. This was particularly evident in the development of like and , where small amounts of tin or depressed the of from about 1085°C to around 950°C or lower, depending on composition. Building on these empirical insights, French chemist François-Marie Raoult conducted systematic experiments in the 1880s on the vapor pressure of solutions, which provided a key link to colligative freezing and melting point depressions. Between 1882 and 1887, Raoult measured the vapor pressure lowering caused by non-volatile solutes in various solvents, establishing that the depression is proportional to the of the solute. For instance, his work with in showed consistent reductions in , leading to the formulation of and its extension to predict boiling and freezing point shifts, including melting point depression in analogous solid-liquid systems. Hints of melting point depression in nanoscale systems emerged in the mid-20th century through advances in . In 1954, Japanese physicist Masao Takagi used to study thin metal films and small particles of metals like tin, lead, and , observing that particles with radii below 20 melted at temperatures significantly lower than their bulk counterparts—for example, tin particles of about 5 radius melted around 100–200°C below the bulk value of 232°C. These early observations demonstrated size-dependent effects, attributing the depression to increased contributions in small particles, paving the way for later nanomaterial .

Key Theoretical Contributions

The foundational theoretical framework for melting point depression in solutions emerged in the late through extensions of laws to phase equilibria. François-Marie Raoult's 1882 work established that the depression of the freezing point (\Delta T_f) in dilute solutions is proportional to the of the solute (x), expressed as \Delta T_f = K_f x, where K_f is the derived from the solvent's of and . This quantitative linked solute concentration directly to the shift in the solid-liquid equilibrium temperature, providing a cornerstone for . Building on Raoult's empirical observations, Jacobus Henricus van't Hoff developed the ideal solution theory in the 1880s, integrating osmotic pressure with phase equilibria to explain colligative effects thermodynamically. Van't Hoff's 1887 formulation treated dilute solutions as analogous to ideal gases, deriving the osmotic pressure \pi = cRT (where c is molar concentration, R is the gas constant, and T is temperature) and connecting it to vapor pressure lowering, which in turn governs freezing point depression via the equality of chemical potentials at equilibrium. This theoretical unification allowed predictions of \Delta T_f from molecular kinetics, emphasizing entropy changes in ideal mixtures without specific solute-solvent interactions. In the 20th century, attention shifted to non-ideal behaviors in alloys and concentrated solutions, where 's regular solution model (1929) accounted for enthalpic contributions to mixing. Hildebrand proposed that in regular solutions, the excess arises solely from a positive \Delta H = \beta x_1 x_2 (with \beta as the interaction parameter and x_1, x_2 as mole fractions), while follows ideal mixing, leading to deviations from and enhanced melting point depression in systems with unfavorable interactions. This model formalized how non-ideal enthalpies alter phase diagrams, particularly in binary alloys, enabling quantitative assessment of eutectic points and limits. For small particles, the introduction of surface energy effects provided a distinct theoretical basis for size-dependent melting point depression. Pawlow's 1909 thermodynamic analysis derived that the melting temperature depression \Delta T scales inversely with particle radius r as \Delta T \propto \frac{\sigma}{r \Delta H_f \rho}, where \sigma is the solid-liquid interfacial energy, \Delta H_f is the latent heat of fusion, and \rho is density; this arises from the excess free energy due to surface curvature in the Gibbs-Thomson relation. Later refinements in the 1970s by Buffat and Borel incorporated experimental validation for nanoparticles, confirming the $1/r dependence while adjusting for liquid skin formation around solid cores. A key conceptual adaptation involved Lindemann's 1910 melting criterion, which posits that melting occurs when the root-mean-square atomic displacement reaches a fraction (approximately 0.1–0.15) of the interatomic distance due to thermal vibrations. Originally for pure , this vibrational instability model was extended to mixtures and alloys, where solute-induced distortions reduce the critical displacement threshold, thereby depressing the in proportion to the disorder or concentration; such adaptations explain deviations in solid solutions beyond purely thermodynamic models.

Measurement Techniques

Experimental Methods

In laboratories, the of solid compounds is commonly determined using the capillary tube method to assess purity through melting-point depression. A small amount of finely ground sample (typically 1–2 mm) is packed into one end of a thin tube (1–2 mm diameter), which is then attached to a or placed in a heating apparatus such as an or electric heating block (e.g., Mel-Temp device). The is raised gradually at a controlled rate (about 1–2 °C/min near the expected ), and the range from when the sample first softens to fully liquefies is recorded using a magnifying or light source. Impure samples exhibit a depressed onset and broader range (often >2 °C) compared to pure compounds (0.5–1 °C range), directly indicating contamination. This straightforward technique is fundamental for routine purity checks in synthesis workflows. Differential scanning calorimetry (DSC) is a widely used technique for quantifying melting point depression in bulk materials such as alloys and solutions, where it records heating curves that reveal shifts in the onset temperature of endothermic melting peaks compared to pure substances. The method achieves high precision, with temperature resolution typically down to 0.1°C, allowing detection of subtle depressions due to solute incorporation or alloying elements. For example, in aluminum-based alloys, DSC identifies the solidus and liquidus temperatures, highlighting eutectic depressions through the shape and position of the melting endotherm. Hot-stage microscopy provides direct visual observation of melting events under controlled heating, particularly useful for alloys and nanoparticles where morphological changes during transitions are evident. Samples are placed on a heated stage within an , and is monitored in real-time as the material softens, flows, or loses crystallinity, enabling qualitative assessment of by comparing observed transition temperatures to bulk references. This technique is especially effective for heterogeneous alloys, where it captures localized behaviors not resolvable by calorimetric methods alone. Thermogravimetric analysis (TGA) coupled with is employed for volatile mixtures, such as solutions or low-melting alloys with evaporative components, to simultaneously track mass loss and heat flow during heating. The integration allows differentiation of melting endotherms from volatilization events, ensuring accurate identification of depressed melting points influenced by volatile solutes. For instance, in organic mixtures, the coupled system reveals how solvent evaporation contributes to apparent depression while isolating true signals. In studies, (TEM) with heating holders tracks individual melting events, providing atomic-scale insights into size-dependent depression. By observing shape changes or liquid-like mobility in real-time under electron beam illumination, researchers quantify depression for particles as small as 10 nm, such as in tin or systems where melting initiates at surface sites. Sample preparation is critical for reliable measurements, with alloys often requiring rapid from the melt to preserve metastable structures and prevent segregation that could mask effects. For solutions, freezing protocols involve controlled cooling rates to form homogeneous eutectic mixtures, ensuring uniform solute distribution before subsequent analysis. These steps minimize artifacts, aligning experimental observations with thermodynamic expectations of colligative lowering.

Data Interpretation

In (DSC) analysis of melting-point depression, baseline correction is essential to isolate the true thermal events from instrumental artifacts. This involves subtracting the signal obtained from an empty reference pan, run under identical conditions to the sample, which accounts for asymmetries in heat flow and sensor imbalances, thereby accurately identifying the onset, peak, and endset of melting transitions. Extrapolation methods, such as constructing linear plots of melting-point depression (ΔT) versus solute concentration (or ), enable the determination of colligative behavior and identification of deviations due to non-ideal interactions. These van't Hoff-style plots, where the slope relates to the and van't Hoff factor, allow extrapolation to zero concentration to obtain the pure solvent's , revealing activity coefficients or effects in the . Error sources like and kinetic can significantly distort measured depression values, as lowers the observed freezing onset during cooling while widens the gap between melting and solidification temperatures. These effects, often quantified by comparing cooling and reheating cycles in successive DSC scans, arise from barriers and can be as large as 150 in nanoscale systems, necessitating multiple cycles to average out kinetic influences and estimate equilibrium conditions. Statistical fitting techniques, particularly least-squares applied to data from multiple runs across compositions, facilitate the construction of accurate diagrams by minimizing residuals between observed and modeled transition temperatures. This approach optimizes parameters for non-ideal mixing models, ensuring robust phase boundaries for eutectic or peritectic systems while accounting for experimental variability. For nanoparticles, where melting peaks in are broadened due to polydispersity, size distribution using log-normal fits explains the observed width by simulating the superposition of size-dependent depression curves, with the directly correlating to peak . This method, informed by complementary techniques like , refines apparent melting temperatures by weighting contributions from the distribution's mean and variance.

Bulk Materials

Alloys and Eutectics

In binary alloy systems, melting-point depression manifests prominently through the formation of eutectic compositions, where the mixture achieves the lowest possible melting temperature compared to the pure components. This occurs at a specific composition where the liquidus curves of the two components intersect in the , resulting in simultaneous solidification of both solid phases upon cooling below the eutectic temperature. For instance, in the lead-tin (Pb-Sn) system, the eutectic composition at approximately 61.9 wt% Sn and 38.1 wt% Pb melts at 183°C, significantly lower than the melting points of pure Pb (327°C) and pure Sn (232°C), enabling applications requiring reduced processing temperatures. The magnitude of this depression and the relative proportions of phases in the two-phase region of a can be quantified using the , which calculates the mass or mole fractions of each phase based on the overall alloy and the tie-line endpoints at a given . This rule illustrates how deviations from the eutectic lead to varying degrees of liquid persistence during solidification, influencing the extent of melting-point lowering; for example, hypoeutectic alloys (with less of the higher-melting component) exhibit a melting range rather than a sharp eutectic point, but still benefit from depressed onset temperatures relative to pure metals. During solidification of non-eutectic alloys, solute segregation at the solid- can induce constitutional , where the ahead of the advancing front becomes compositionally enriched and thus cools below its equilibrium freezing point, promoting in the planar . This leads to dendritic patterns, as small perturbations amplify into branching structures that trap solute in interdendritic regions, resulting in uneven local melting-point depression across the microstructure and potential microsegregation effects that alter overall thermal behavior. A notable example is the gold-silicon (Au-Si) system, with a eutectic at 3 wt% Si (19 at% Si) melting at 363°C—well below pure Au (1064°C) and Si (1414°C)—which is exploited in for in micro-electro-mechanical systems () due to its ability to form strong, hermetic seals at moderate temperatures without damaging sensitive components. Industrially, eutectic alloys are pivotal in solders, such as the Pb-Sn composition for electronic assembly, where the depressed facilitates low-temperature joining processes that minimize on substrates and devices. Similarly, in alloys like aluminum-silicon systems, eutectic formation enables improved fluidity and reduced pouring temperatures, enhancing mold filling and reducing defects in applications ranging from automotive parts to components.

Solutions and Colligative Properties

Melting point depression manifests as freezing point depression in liquid solutions, a that depends on the number of solute particles rather than their identity. In molecular mixtures and nonelectrolyte solutions, the addition of a solute lowers the freezing point of the by interfering with the formation of the pure 's solid phase. This effect is quantified by the formula \Delta T_f = K_f \cdot m \cdot i, where \Delta T_f is the in degrees , K_f is the of the (1.86 °C/m for ), m is the of the solute, and i is the van't Hoff factor representing the number of particles per solute molecule (i=1 for nonelectrolytes). A practical application of this property is in antifreeze formulations, where ethylene glycol is mixed with water to prevent ice formation in vehicle cooling systems. A 70% by volume ethylene glycol-water mixture depresses the freezing point to approximately -50°C, protecting engines from damage in subzero conditions by ensuring the solution remains liquid. Urea, a nonelectrolyte solute, is employed in water-based de-icing solutions for applications like airport runways, where a 20% concentration by weight yields a freezing point depression of up to 10°C, facilitating ice removal without excessive environmental impact compared to salts. In solutions, such as those containing salts, the van't Hoff factor i exceeds 1 due to into s, leading to greater than in nonelectrolyte solutions at equivalent ; however, ion pairing reduces the effective i below the ideal value, particularly in concentrated solutions, as ions associate to form neutral pairs that behave like single particles. At high solute concentrations, the colligative behavior deviates due to non-ideality, where solute-solvent and solute-solute interactions alter activity coefficients, causing the observed \Delta T_f to differ from predictions of the linear formula and often resulting in less depression than expected.

Nanomaterials

Size-Dependent Effects

In , reducing the particle size to the nanoscale significantly amplifies melting point depression compared to materials, primarily due to the increased surface-to-volume . As particle decreases, the proportion of surface atoms rises sharply, enhancing the contribution of surface to the overall of . This leads to a size-dependent depression where the change in melting temperature, ΔT, is inversely proportional to the diameter d (ΔT ∝ 1/d), making surface effects dominant over energies. This phenomenon was first theoretically described by Pawlow in through a thermodynamic model equating the chemical potentials of solid and liquid phases for spherical particles, yielding the size-dependent melting temperature: T_m(d) = T_m(\infty) \left(1 - \frac{\alpha}{d}\right) where T_m(\infty) is the bulk melting temperature and \alpha is a material-specific constant typically ranging from 1 to 2 nm for metals, reflecting the ratio of surface energy differences to latent heat and density. Experimental observations confirm this scaling, as seen in gold nanoparticles where 2 nm particles melt at approximately 300°C, far below the bulk value of 1064°C, due to the overwhelming influence. In semiconductors, additional quantum confinement effects contribute subtly; the widening bandgap with decreasing size stabilizes electronic states but indirectly destabilizes the by altering vibrational modes and , exacerbating melting point depression in nanocrystals like . For nanoparticles, representative measurements show depressions of 50–100°C at scales around 10 nm, aligning with the enhanced surface contributions while highlighting the material's sensitivity to nanoscale geometry.

Shape and Substrate Influences

The of nanoparticles plays a crucial role in their behavior, leading to anisotropic effects that deviate from isotropic spherical models. Rod-shaped nanoparticles, for example, exhibit greater melting point depression compared to spheres of equivalent volume due to their elongated geometry and higher surface-to-volume ratio, which amplifies contributions. In rods, often initiates at the tips, where local curvature is highest, consistent with the Gibbs-Thomson relation that correlates temperature inversely with . Substrate interactions further modulate this depression through interfacial effects, particularly influenced by the wetting angle between the , its melt, and the support. Hydrophilic substrates, with low s promoting strong , enhance liquid phase pinning and spreading, resulting in additional melting point reductions of 20–50 °C for small nanoparticles by stabilizing thinner solid phases. In contrast, hydrophobic or weakly interacting supports minimize such effects, preserving higher melting temperatures. This wetting dependence has been demonstrated in simulations of bcc-metal nanoparticles, where contact angle variations directly correlate with the extent of thermal depression. In supported catalytic systems, this depressed melting temperature enhances dissolution rates of nanoparticle components into surrounding media or electrolytes, promoting atomic-level restructuring and improved reactivity without full coalescence. Such effects are particularly beneficial for oxygen reduction reaction catalysts, where shape-dependent dissolution at lowered temperatures optimizes active site exposure. Experimental atomic force microscopy (AFM) imaging has captured shape-induced melting fronts in nanoparticles, revealing preferential initiation at high-curvature features like rod tips, with progressive surface softening visualized through nanoscale topography changes during controlled heating. These observations complement broader size-dependent trends by highlighting geometry-specific dynamics.

Theoretical Models

Classical Approaches

The Gibbs-Thomson effect provides a foundational thermodynamic description of melting point depression arising from the of the solid-liquid in finite-sized systems. This effect originates from the increase in due to contributions, analogous to the elevation of over curved liquid surfaces, and applies to the between a solid particle and its melt. For a spherical particle, the depression in melting temperature \Delta T is expressed as \Delta T = \frac{2 \gamma T_m}{\rho \Delta H_f r}, where \gamma is the solid-liquid interfacial energy, T_m is the bulk melting temperature, \rho is the solid density, \Delta H_f is the bulk latent heat of fusion per unit mass, and r is the particle radius. This continuum model predicts a linear inverse dependence on radius, with depressions becoming significant for particles below 100 nm, and has been validated experimentally for metals like gold and tin. The Lindemann criterion offers an alternative classical perspective, viewing as the point where thermal vibrations destabilize the crystal . It stipulates that initiates when the root-mean-square atomic displacement reaches approximately 10-15% of the nearest-neighbor interatomic distance, corresponding to a loss of long-range order. In impure systems, such as dilute alloys, introduce local distortions that enhance vibrational amplitudes at lower temperatures, thereby depressing the through increased disorder. This mechanism explains solute-induced depressions in metals, where even small concentrations (e.g., 1-5 at.%) can lower T_m by tens of degrees , consistent with empirical observations in binary systems. For multicomponent alloys, regular solution theory quantifies melting point depression by incorporating non-ideal mixing effects into the phase equilibrium. Assuming random atomic distribution without long-range order, the theory posits an excess of mixing \Delta G_{ex} = \Omega x (1 - x), where \Omega is the regular solution interaction parameter (positive for endothermic mixing) and x is the solute . This excess term shifts the common tangent construction in the diagram, lowering the and liquidus temperatures relative to solutions and predicting depression depths proportional to \Omega and , as seen in systems like Cu-Ni where \Omega \approx 15 kJ/mol yields 20-50 K reductions at equiatomic compositions. These models extend to colloidal emulsions, where droplet confinement mimics curvature. In oil-in-water emulsions of alkanes or fats, the Gibbs-Thomson effect causes surface layers to melt at temperatures 5-10°C below the bulk value for droplets of 1-10 \mum radius, facilitating partial coalescence and influencing stability in foods like . Despite their utility for bulk alloys and larger colloids, classical approaches like Gibbs-Thomson and Lindemann exhibit limitations for nanoparticles smaller than 5 nm, where assumptions break down and atomic structures dominate, leading to deviations up to 50% from predicted depressions.

Nanoparticle-Specific Mechanisms

In nanoparticle-specific mechanisms of melting point depression, the emergence of core-shell structures plays a central role, particularly in metallic and covalent systems. Surface atoms in nanoparticles possess reduced coordination numbers relative to the , resulting in weaker overall binding and initiating premature of an outer liquid-like while the core remains solid. This two-stage process, where the melts at temperatures significantly below the due to the high surface-to-volume ratio, has been elucidated through atomic-scale simulations of bimetallic nanoparticles, highlighting how the shell's instability drives the subsequent core . Bond contraction at surfaces represents another distinctive mechanism contributing to depression. The fewer neighboring atoms at the surface lead to shorter interatomic lengths, enhancing the strength of individual bonds but generating on the surface and tensile in the subsurface layers. This destabilizes the core , lowering the barrier for and amplifying the size-dependent depression, as evidenced by experimental measurements on metallic nanoparticles like and tin, where bond lengths contract by up to several percent for particles below 10 nm. Semiconductor nanoparticles, such as , demonstrate particularly pronounced deviations from classical predictions. For instance, silicon nanoparticles below 5 nm have been reported to melt at temperatures as low as 200°C relative to the bulk value of 1414°C, attributed to the combined effects of surface premelting and quantum confinement that further reduce stability beyond macroscopic models. Assessing these mechanisms through empirical fitting underscores the limitations of isolated theories, as single models often fail to reconcile experimental data across particle sizes and materials. Hybrid approaches, integrating core-shell dynamics, bond strain, and contributions, provide superior accuracy, with fitting errors reduced by factors of 2–5 compared to classical or purely thermodynamic models when validated against data for metals and semiconductors. A notable post-2020 advancement involves frameworks trained on datasets to predict size-depression curves for melting points. These models, such as Gaussian process regressions applied to and aluminum systems, achieve predictions within 5–10% of experimental values, enabling efficient exploration of composition effects without full-scale simulations.

Liquid Drop Model

The liquid drop model (LDM) for melting-point depression in nanoparticles draws a direct analogy to the nuclear liquid drop model, treating metallic nanoparticles or clusters as deformable charged droplets where short-range metallic bonds mimic nuclear forces. In this framework, the melting transition occurs when thermal energy disrupts the balance between , which favors a compact spherical , and repulsion from excess charge, which promotes instability and expansion. This balance determines the energy barrier for the phase change, analogous to the barrier in nuclei, but adapted here to describe lattice —the breakup of the atomic lattice during melting rather than binary splitting. The key equation for the melting energy E_m in this model is given by E_m = 4\pi r^2 \sigma + \frac{3}{5} \frac{(Z e)^2}{r}, where r is the cluster radius, \sigma is the surface tension, Z is the effective charge number, and e is the . The first term represents the surface energy cost of forming the interface during melting, while the second term accounts for the electrostatic self-repulsion in the charged droplet, adapted from to metallic where the uniform charge within the cluster amplifies instability at elevated temperatures. This formulation predicts that smaller clusters, with higher surface-to-volume ratios and intensified effects for fixed charge, exhibit greater melting-point depression. For metallic clusters, the model forecasts a depression scaling as \Delta T \propto 1/r^{2/3}, arising from the dominant surface contribution to the per-atom energy difference between solid and liquid phases, modulated by charge effects in small systems. This scaling captures the enhanced instability in nanoscale droplets compared to bulk materials. When applied to alkali metals, the LDM accurately reproduces experimental melting temperatures for sodium (Na) clusters down to approximately 100 atoms, where depressions of up to 28% relative to the bulk value of 371 K are observed, aligning with calorimetric measurements of latent heat reductions. Molecular dynamics simulations validate the model's depiction of drop-like instability, showing that surface atoms in heated metallic clusters undergo premature disordering and collective fluctuations resembling a , with the overall structure destabilizing in a manner consistent with macroscopic droplet dynamics under thermal and electrostatic stresses.

Liquid Shell Nucleation Model

The Liquid Shell Nucleation Model posits that melting in nanoparticles begins with surface premelting, where a thin layer, approximately 1 nm thick, forms on the core at temperatures below the bulk due to the reduced of surface atoms compared to those in the interior. This reduced coordination elevates the of surface atoms, facilitating their transition to a liquid-like state and initiating the premelting process. The model, originally proposed by Reiss and in 1948, emphasizes that this quasi-liquid shell remains stable as the temperature approaches the effective of the nanoparticle. The formation of this liquid shell is governed by a nucleation barrier derived from classical nucleation theory, adapted for surface-initiated melting: \Delta G_\text{nuc} = -\frac{4}{3} \pi r^3 \Delta \mu + 4 \pi r^2 \gamma_\text{sl} Here, r represents the radius of the embryonic liquid nucleus, \Delta \mu is the bulk free energy difference per unit volume driving the solid-to-liquid transition (which becomes more favorable below the bulk melting temperature), and \gamma_\text{sl} is the solid-liquid interfacial energy. For nanoparticles, the high surface-to-volume ratio reduces this barrier, promoting easier nucleation of the liquid phase at the surface and contributing to the observed melting point depression. Following , the liquid shell grows by thickening inward, progressively consuming the solid core until complete melting occurs through core collapse. This mechanism aligns with experimental observations from (TEM) studies on silver nanoparticles, where the model accurately predicts a melting point depression of about 150°C for particles around 10 nm in diameter. However, the model overestimates the extent of melting point depression in covalent semiconductors like , where stronger directional bonding limits surface premelting compared to metallic systems.

Liquid Nucleation and Growth Model

The liquid nucleation and growth model (LNG) conceptualizes the of nanoparticles as a heterogeneous process initiating at the surface, where a nucleus forms due to reduced atomic coordination and elevated , followed by the inward propagation of the liquid-solid interface. This model adapts to nanoscale dimensions, emphasizing that the high surface-to-volume ratio enhances the thermodynamic driving force for . Unlike bulk materials, nanoparticles exhibit surface premelting, allowing the phase to nucleate more readily at defects or edges before expanding volumetrically. The for the in this model is derived from as r_c = \frac{2\gamma}{\Delta G_v}, where \gamma is the solid-liquid interfacial energy and \Delta G_v is the change per unit volume between the solid and phases. In nanoparticles, \Delta G_v is amplified by effects and surface contributions, yielding a smaller r_c than in bulk systems, which facilitates at lower temperatures and contributes to melting point depression. This adaptation highlights how nanoscale confinement lowers the energy barrier for , enabling surface-initiated melting even below the bulk melting temperature. Once nucleated, the liquid phase grows through the nanoparticle via the motion of the solid-liquid interface, governed by kinetic considerations. The interface velocity is expressed as v = M \left(1 - \exp\left(\frac{\Delta G}{RT}\right)\right), where M is the atomic mobility at the interface, \Delta G is the chemical driving force for growth, R is the gas constant, and T is the temperature. As temperature rises, the exponential term decreases, accelerating v and promoting rapid inward propagation, which amplifies the effective melting point depression observed experimentally. This kinetic acceleration distinguishes the LNG model from equilibrium-based approaches, providing a dynamic explanation for the temperature range over which melting occurs. The LNG model accounts for thermal in cyclic heating and cooling of nanoparticles, where proceeds via surface but solidification requires homogeneous within the supercooled liquid, leading to undercooling. For nanoparticles, this manifests as in reheating cycles, with undercooling on the order of 50 K during solidification, as the liquid resists recrystallization until a critical is reached. This behavior underscores the model's utility in interpreting repeated thermal cycling effects in . Molecular dynamics (MD) simulations integrated with the LNG framework reveal that liquid growth proceeds in three dimensions from surface defects, such as vacancies or edges, confirming the heterogeneous initiation and volumetric expansion predicted by the model. These simulations demonstrate defect-driven sites facilitating 3D propagation, with the liquid phase enveloping the core progressively, aligning with experimental observations of size-dependent melting. In practical applications, the LNG model aids in predicting the thermal stability of catalysts under operational heating, such as in iron-based systems for synthesis, where premature surface melting could degrade performance. By quantifying the onset of liquid propagation, the model informs design strategies to maintain structural integrity at elevated temperatures, enhancing catalyst in high-heat environments.

Bond-Order-Length-Strength Model

The Bond-Order-Length-Strength (BOLS) model offers an atomistic explanation for melting point depression in nanomaterials by correlating atomic coordination number with interatomic bond length and strength, emphasizing the role of surface undercoordination. Developed by C.Q. Sun and collaborators, the model describes how reduced coordination at the surface of nanoparticles leads to local structural and energetic changes that lower the overall melting temperature compared to the bulk material. This approach bridges quantum-chemical bond analysis with thermodynamic properties, providing a framework for both metallic and non-metallic systems. A central feature of the BOLS model is the core-shell bond contraction, where surface atoms with fewer nearest neighbors form shorter and stronger s than those in the , typically by 10-20%, while the experiences corresponding tensile . This arises from the spontaneous relaxation of undercoordinated atoms to minimize , quantified by the relation for d_z = C_z^m d_b, where d_z is the length at coordination number z, C_z is the given by C_z = \left[ 1 + \exp\left( \frac{z_z - z_b}{8 z_z} \right) \right]^{-1}, m is the scaling exponent (approximately for covalent bonds), and d_b is the . The resulting bond strengthening elevates local cohesive at the surface but reduces the average cohesive across the nanoparticle as size decreases, due to the increasing surface-to-volume ratio. The melting criterion in the BOLS model follows from the Lindemann instability, positing that the melting temperature T_m is proportional to the ratio of average to vibrational , T_m \propto E_b / \nu. Bond contraction raises \nu (blueshifting vibrations) but the dominant effect is the size-induced drop in average E_b, leading to expressed as T_m(d) / T_{bulk} = 1 - (C_n / C_{bulk})^\beta, where C_n is the effective of the (dependent on diameter d), C_{bulk} is the coordination, and \beta (typically 0.5-1) accounts for bond-order deficiency. A derived empirical form for the relative is \Delta T / T_{bulk} = 1 - \exp\left[ -(\tau / d) \right], with characteristic length \tau \approx 0.5 reflecting the surface shell thickness. Extensions of the BOLS model to oxide semiconductors (e.g., ZnO, TiO₂) and organic nanomaterials address limitations in prior models for non-metallic systems, by incorporating covalent bond-order variations to explain observed depressions in cohesive energy and melting points. For instance, applications to GaAs and InP nanocrystals correlate band gap expansion with melting depression through coordination-resolved bond changes, filling theoretical gaps for systems with directional bonding.

References

  1. [1]
    6.1C: Melting Point Theory - Chemistry LibreTexts
    Apr 7, 2022 · Melting point depression is the result of different changes in entropy when melting a pure and impure solid. As solids are restricted in atomic ...
  2. [2]
    [PDF] Melting point determination
    A mixture of very small amounts of miscible impurities will produce a depression of the melting point and an increase in the melting point range.
  3. [3]
    5.3: MELTING POINT ANALYSIS- IDENTITY AND PURITY
    Jun 20, 2021 · The presence of impurities will influence the melting point of a compound, leading to wider and depressed melting point ranges.
  4. [4]
    [PDF] Using Melting Point to Determine Purity of Crystalline Solids
    May 18, 2009 · Less heat corresponds to a lower temperature; thus, an impure solid melts at a lower temperature than the same solid with no impurities present.<|control11|><|separator|>
  5. [5]
    12.5 Colligative Properties – Chemistry Fundamentals
    Colligative properties include vapor pressure lowering, boiling point elevation, freezing point depression, and osmotic pressure, depending on total solute ...
  6. [6]
    Colligative Properties
    A colligative property if it depends only on the ratio of the number of particles of solute and solvent in the solution, not the identity of the solute.Missing: scientific | Show results with:scientific<|control11|><|separator|>
  7. [7]
    Binary Eutectic Phase Diagrams
    The binary eutectic phase diagram explains the chemical behavior of two immiscible (unmixable) crystals from a completely miscible (mixable) melt, ...
  8. [8]
    [PDF] Effect of Multiple Alkanethiol Ligands on Solubility and Sintering ...
    May 25, 2010 · Due to the nanoscale geometries, the gold particles exhibit vastly reduced melting temperatures from the bulk temperature of 1064 °C. Pawlow22 ...
  9. [9]
    [PDF] Controlling the Structure of Two-dimensional Nanoparticle ...
    depression, which melt at lower temperatures than the same bulk material (Figure 1-10). For example, bulk gold has a melting point of around 1064. ◦. C (60) ...
  10. [10]
  11. [11]
    Gibbs free energy and spontaneity (article) - Khan Academy
    Thus, we see that at − 10 ∘ C ‍ the Gibbs free energy change Δ G ‍ is positive for the melting of water. Therefore, we would predict that the reaction is not ...
  12. [12]
  13. [13]
    Mixing and electronic entropy contributions to thermal energy ...
    Jul 11, 2017 · Melting of crystalline solids is associated with an increase in entropy due to an increase in configurational, rotational, and other degrees ...
  14. [14]
    The pre-history of cryoscopy: what was done before raoult? - SciELO
    Blagden undertook in 1788 the study of the effect of different dissolved inorganic substances in the lowering of freezing point of water9 as a continuation of ...
  15. [15]
    Paper, 'Experiments on the effect of various substances in lowering ...
    Paper, 'Experiments on the effect of various substances in lowering the point of congelation in water' by Charles Blagden. Reference number: L&P/9/85.Missing: freezing salt solutions
  16. [16]
    [PDF] Metallography and Microstructure of Ancient and Historic Metals
    This volume is an attempt to provide a measured amount of infor mation regarding the techniques of metallogra phy as they apply to ancient and historic metals.
  17. [17]
    Raoult vapor pressure depression - Le Moyne
    François-Marie Raoult investigated the depression of vapor pressure of a variety of solvents with many different non-volatile solutes dissolved in them.
  18. [18]
    Electron-Diffraction Study of Liquid-Solid Transition of Thin Metal Films
    The observed melting points are found to be lower than those of bulk metals. Anticipating this effect to be attributed to the small size of the crystal, the ...Missing: microscopy particles
  19. [19]
    The size dependence of the melting point of small particles of tin
    Aug 6, 2025 · The size dependence of the melting point of tin has been studied by means of transmission electron diffraction and microscopy.
  20. [20]
    Raoult freezing point depression - Le Moyne
    François-Marie Raoult investigated the depression of the freezing point of a variety of solvents with many different solutes dissolved in them.
  21. [21]
    van 't Hoff and Osmosis - chemteam.info
    With solutions we deal with the so-called osmotic pressure, where with gases we are concerned with the ordinary elastic pressure.Missing: phase 1880s primary source
  22. [22]
    Size-dependent melting of nanoparticles: Hundred years of ...
    Aug 9, 2025 · Pawlow deduced an expression for the size-dependent melting temperature of small particles based on the thermodynamic model which was then ...Missing: Buffer 1960s
  23. [23]
    [PDF] DTA and Heat-flux DSC Measurements of Alloy Melting and Freezing
    This document is focused on differential thermal analysis (DTA)* and heat- flux differential scanning calorimetry (HF-DSC)† of metals and alloys. A ther ...
  24. [24]
    Differential Scanning Calorimetry (DSC) and Thermodynamic ...
    Jul 31, 2017 · The melting of these two alloys was investigated using differential scanning calorimetry (DSC) and thermodynamic prediction.
  25. [25]
    Hot stage microscopy and its applications in pharmaceutical ... - PMC
    Jun 16, 2020 · Hot stage microscopy (HSM) is a thermal analysis technique that combines the best properties of thermal analysis and microscopy.
  26. [26]
    Hot Stage Microscopy - Microtrace
    By varying the temperature, the melting point of a substance can be determined. The hot stage can also be used to study polymorphic transformations, phase ...Missing: depression alloys
  27. [27]
    Hot Stage Microscopy | Thermo Fisher Scientific - US
    Mar 13, 2020 · Hot stage microscopy adds an SEM heating stage to FIB SEM, enabling the in situ study of materials as they recrystallize, melt, deform, ...Missing: depression | Show results with:depression
  28. [28]
    [PDF] Thermogravimetry (TG) & Differential Scanning Calorimetry (DSC)
    The complimentary information obtained allows differentiation between endothermic and exothermic events which have no associated weight loss (e.g., melting and ...
  29. [29]
    Coupled and Simultaneous Thermal Analysis Techniques in the ...
    May 25, 2023 · DSC–photovisual allows direct observation of thermal transformations occurring in the sample during heating or cooling under DSC conditions. DSC ...
  30. [30]
    [PDF] Thermogravimetric Analysis TG 309 Libra®
    When mixtures or blends are measured at reduced pressure, boiling point depression can be realized for volatiles (e.g., solvents, plasticizers). This leads.
  31. [31]
    Direct Imaging of Surface Melting on a Single Sn Nanoparticle
    Jul 7, 2023 · We used Sn nanoparticles (NPs), which were heated inside a transmission electron microscope (TEM) to observe the kinetics of melting and ...
  32. [32]
    In Situ Transmission Electron Microscopy Investigation of Melting ...
    Nov 30, 2021 · Sambles J.R. An electron microscope study of evaporating gold particles: The kelvin equation for liquid gold and the lowering of the melting ...Missing: 1950s | Show results with:1950s
  33. [33]
    Influence of quenching on the martensitic transformation ...
    The aim of the present work is to reveal the origin of the lowering of martensitic transformation temperatures in quenched Au–47.5at.%Cd–Ag alloys. Section ...
  34. [34]
    A Low-Cost and Simple Demonstration of Freezing Point ...
    Sep 8, 2022 · A laboratory activity was developed to teach freezing point depression and colligative properties to introductory-level chemistry students.
  35. [35]
    Freezing Point Depression - Chemistry LibreTexts
    Feb 26, 2023 · Freezing point depression is a colligative property observed in solutions that results from the introduction of solute molecules to a solvent.
  36. [36]
    [PDF] Differential Scanning Calorimetry (DSC) Theory and Applications
    The heat flow baseline is usually curved and has large slope and offset. Loss of sensitivity as a result of curvature in the baseline. • The heating rate of the ...
  37. [37]
    Freezing point depression in model Lennard-Jones solutions
    Here we investigate the basic principles of solute-induced freezing point depression by computing the melting temperature of a Lennard-Jones fluid with low ...
  38. [38]
    A large thermal hysteresis between melting and solidification ...
    A very large thermal hysteresis loop width of 213K that is, the depression of the melting point and the increase of supercooling of Pb particles with a diameter ...
  39. [39]
    Non-linear least squares analysis of phase diagrams for non-ideal ...
    A computer program for non-linear least squares minimization has been applied to construct temperature-composition phase diagrams for several binary systems ...
  40. [40]
    Effect of grain size on the melting point of confined thin aluminum films
    Oct 23, 2014 · The modeled nanocrystalline structures had a log-normal distribution of grains diameter and a random distribution of grain orientations. All ...
  41. [41]
    Lead-free Solders in Microelectronics - ScienceDirect.com
    With a melting eutectic temperature of 183°C, the Sn–Pb binary system allows soldering conditions that are compatible with most substrate materials and devices.
  42. [42]
    [PDF] General Discussion of Phase Diagrams
    The eutectic composition is that combination of com- ponents in a simple system having the lowest melting temperature of any ratio of the components and is.
  43. [43]
    12.7: The Lever Rule - Engineering LibreTexts
    Nov 26, 2020 · If an alloy consists of more than one phase, the amount of each phase present can be found by applying the lever rule to the phase diagram.Lever rule applied to a binary... · Point 3
  44. [44]
    The lever rule
    If an alloy consists of more than one phase, the amount of each phase present can be found by applying the lever rule to the phase diagram.
  45. [45]
    Constitutional Supercooling - an overview | ScienceDirect Topics
    Each dendrite grows in a random direction until finally the arms of the dendrites are filled, and further growth is obstructed by the neighboring dendrite. As a ...
  46. [46]
    Evolution of dendritic patterns during alloy solidification
    The physical origin of this instability was described in 1953 by Rutter and Chalmers (1) and by Tiller et al. (2) as “constitutional supercooling.” Because the ...
  47. [47]
    Investigation of Au/Si Eutectic Wafer Bonding for MEMS ...
    May 15, 2017 · Au/Si eutectic bonding is considered to BE a promising technology for creating 3D structures and hermetic packaging in micro-electro-mechanical ...
  48. [48]
    [PDF] A Robust Gold-Silicon Eutectic Wafer Bonding Technology for ...
    Gold-silicon eutectic formation occurs at 363°C for 19 atomic % Si, as shown in the phase diagram in Fig. 1. The eutectic can be used to bond two wafers, or be ...
  49. [49]
    Eutectic Point: Basics and Uses - Stanford Advanced Materials
    Jul 24, 2025 · Eutectic systems are widely used in soldering, where low melting points are advantageous. Additionally, they play a critical role in casting ...
  50. [50]
    Applications and Recent Advances of Low-Temperature ... - NIH
    This paper reviews the wettability and preparation methods of low-temperature multicomponent solders, and concludes the effect of different metallic elements ...
  51. [51]
    Freezing Point Depression
    The freezing point of a solution is less than the freezing point of the pure solvent. This means that a solution must be cooled to a lower temperature than the ...
  52. [52]
    [PDF] 1 ∆T = KF·m Lab #12: Freezing Point Depression
    In this experiment we will use the freezing point depression of lauric acid to determine the molar mass of the solute benzoic acid. Lab #12: Freezing Point ...
  53. [53]
    Ethylene Glycol Heat-Transfer Fluid Properties: Density, Data & Charts
    Due to possible slush creation, ethylene glycol and water solutions should not be used in conditions close to freezing points. Water with ethylene glycol - ...
  54. [54]
    [PDF] Experimental phase diagram of the urea-water binary system - HAL
    Dec 11, 2018 · Since the theoretical freezing point of water is 0 °C, we can deduce the influence of the cooling stage on the freezing point obtained ...
  55. [55]
    13.9: Solutions of Electrolytes - Chemistry LibreTexts
    Jul 12, 2023 · The van't Hoff factor is therefore a measure of a deviation from ideal behavior. The lower the van 't Hoff factor, the greater the deviation ...
  56. [56]
    Freezing Point Depression - chemteam.info
    Example #1: Pure benzene freezes at 5.50 °C. A solution prepared by dissolving 0.450 g of an uknown substance in 27.3 g of benzene is found to freeze at 4.18 °C ...
  57. [57]
    13.6: Colligative Properties- Freezing Point Depression, Boiling ...
    Aug 14, 2020 · Freezing point depression depends on the total number of dissolved nonvolatile solute particles, just as with boiling point elevation. Thus an ...
  58. [58]
    Freezing point depression of salt aqueous solutions using the ...
    Apr 4, 2022 · TIP5P68 and/or TIP4P/Ice69 could also be considered as promising models for freezing point depression as their melting points are quite close ...
  59. [59]
    [PDF] Size-dependent melting of nanoparticles: Hundred years of ...
    However, Pawlow in 1909 developed a thermodynamic model [1], that predicts a melting point depression of nanoparticles and the variation is linear with the ...
  60. [60]
    Size-dependent melting point depression of nanostructures: Nanocalorimetric measurements
    ### Summary of Melting Point Depression for Indium Nanoparticles
  61. [61]
    Melting in Semiconductor Nanocrystals - Science
    Temperature-dependent electron diffraction studies on nanocrystals of CdS show a large depression in the melting temperature with decreasing size.
  62. [62]
  63. [63]
  64. [64]
  65. [65]
    Thermodynamic analysis and modification of Gibbs–Thomson ...
    The melting point depression for all metals listed in Table 1 was calculated using Eq. (7), i.e., the modified Gibbs–Thomson equation or Lee's model, where the ...Missing: principles | Show results with:principles
  66. [66]
    The Gibbs−Thomson Equation and the Solid−Liquid Interface
    The Gibbs−Thomson equation (GTE) is valid in the context of determining interfacial surface energies.Missing: depression seminal
  67. [67]
    Size-Dependent Melting Behavior of Colloidal In, Sn and Bi ... - Nature
    Nov 17, 2015 · Microscopic observation of the solidification of small metal droplets J. Appl. Phys. 21, 804–810 (1950). Article ADS Google Scholar.
  68. [68]
    Generalisation of the Lindemann criterion for disordered mixed ...
    Aug 10, 2025 · The well known Lindemann criterion of melting is generalised for binary and ternary systems with full solubility in both the liquid and the ...Missing: depression | Show results with:depression
  69. [69]
    Molecular dynamics simulation of the melting processes of core ...
    Our simulation provides microscopic insights into the melting behaviours of existing and proposed core–shell nanoparticles that would be highly beneficial ...
  70. [70]
    The lattice contraction of nanometre-sized Sn and Bi particles ...
    The observed lattice contraction of nanoparticles is attributed to both the surface bond that contracts with coordination number reduction of surface atoms.<|separator|>
  71. [71]
    Modeling the effects of size and surface broken bond on surface ...
    Oct 31, 2017 · It is found that reducing the number of surface broken-bonds will slow down the decrease of γ(D). For facets of nanocrystals with different ...<|separator|>
  72. [72]
  73. [73]
    Models of size-dependent nanoparticle melting tested on gold
    Aug 7, 2025 · There is a spectrum of models to calculate the variation of melting point depression of the metals and compounds reported in the literature such ...
  74. [74]
    Accurate thermodynamic relations of the melting temperature of ...
    May 1, 2017 · ... < 5 nm. With the same D values, the depression of melting temperature takes the following sequence: ΔTm (sphere) > ΔTm (icosahedron) > ΔTm ...
  75. [75]
    Data-driven simulation and characterisation of gold nanoparticle ...
    Oct 18, 2021 · We characterize the melting mechanism in gold NPs of size 1–6 nm, and predict the melting temperatures in good agreement with experimental data ...
  76. [76]
    Machine-learnt potential highlights melting and freezing of ...
    Apr 8, 2025 · We study the complete thermodynamical cycle, melting/freezing, of Al nanoparticles with various initial geometries between 102 and 104 atoms. We ...
  77. [77]
    Metal-cluster fission and the liquid-drop model | Phys. Rev. A
    Dec 1, 1992 · Metal-cluster fission and the liquid-drop model. Winston A. Saunders ... W. A. Saunders and S. Fedrigo, Chem. Phys. Lett. 156, 14 (1989) ...
  78. [78]
    Melting of Pd clusters and nanowires: A comparison study using ...
    Oct 19, 2005 · Melting point depressions in both systems agree better with a liquid-drop model than with Pawlow's thermodynamic model. Figure 1; Figure 2
  79. [79]
    Size-dependent melting modes and behaviors of Ag nanoparticles
    The size-dependent melting behaviors and mechanisms of Ag nanoparticles (NPs) with diameters of 3.5–16 nm were investigated by molecular dynamics (MD).
  80. [80]
  81. [81]
  82. [82]
    Review and modelling of Melting Point Depression in Metallic Nano ...
    greater for nanoparticles with r <5 nm. Additionally, a variety of models ... Size and Shape Dependent Melting Point Depression of Al, Ag, Au, and Pb ...
  83. [83]
    Correlation between the band gap expansion and melting ...
    Sep 23, 2015 · However, experimental studies and atomic simulations reveal that nano-semiconductors have lower thermal stability than their bulk counterparts.