Fact-checked by Grok 2 weeks ago

Glass transition

The glass transition is the nonequilibrium process by which an supercooled transforms into a nonequilibrium () upon cooling, or vice versa upon heating, marked by a dramatic slowdown in structural relaxation times without any change in the average or molecular . This transition occurs in amorphous materials, including polymers, metallic glasses, and inorganic oxides, and is characterized by a gradual shift from a rigid, brittle glassy state below the glass transition (Tg) to a more viscous, rubbery state above it. Unlike crystalline , the glass transition involves no or abrupt volume change, but rather a kinetic arrest where the material's reaches approximately 1012 Pa·s (or 1013 poise). This viscosity-based definition is phenomenological, and its relationship to a true thermodynamic glass transition—if such a thing even exists—remains unclear, as the glass transition is debated between kinetic and thermodynamic interpretations. The underlying physics of the glass transition remains a central puzzle in condensed matter science, with two competing theoretical frameworks: thermodynamic approaches, which propose an "ideal" transition driven by a rapid loss of configurational entropy leading to a Kauzmann entropy crisis, and kinetic models, which emphasize dynamic heterogeneity and the loss of ergodicity due to increasingly cooperative relaxation processes. Phenomenological theories, such as the Vogel-Fulcher-Tammann equation for viscosity and mode-coupling theory for the dynamics near Tg, capture universal features like the super-Arrhenius increase in relaxation times and the emergence of a dynamic crossover temperature Td above Tg. These aspects highlight the transition's universality across disparate glass-forming systems, from fragile organic liquids to strong network formers like silica. In practice, Tg governs the mechanical and thermal properties of amorphous materials, setting limits on their processing temperatures, dimensional stability, and performance in applications ranging from structural polymers to pharmaceutical formulations. For instance, in polymers, Tg influences and , with values typically ranging from -100°C for flexible elastomers to over 300°C for high-performance plastics, modulated by factors like chain stiffness, intermolecular interactions, and cooling rate. The transition's sensitivity to thermal history—manifesting as structural relaxation and aging effects—further complicates , underscoring the need for precise control in .

Fundamentals

Characteristics

The glass transition represents a gradual and reversible transformation in amorphous materials, shifting from a hard, brittle glassy to a viscous supercooled upon heating. This process occurs without the absorption or release of and lacks an abrupt change in volume, distinguishing it as a kinetic driven by the freezing of molecular configurations rather than a shift. Across the transition, mechanical properties undergo a marked change: below the glass transition T_g, the material displays high rigidity and , typically on the order of gigapascals, rendering it brittle under . Above T_g, it adopts a rubbery or viscous character with significantly reduced stiffness, often dropping to below 1 in storage , enabling greater flexibility and deformation. The kinetic essence of the glass transition manifests in its sensitivity to experimental conditions, particularly the cooling or heating rate, which influences the point at which structural relaxation halts or resumes. Standard measurements, such as those via , employ rates of 10 K/min to capture this behavior reliably. Unlike , which disrupts long-range crystalline and involves a discontinuous volume expansion with , the glass transition pertains exclusively to amorphous structures devoid of such , emphasizing a slowdown in dynamics over structural reorganization. Early observations of softening behavior in undercooled liquids date to the work of Gustav Tammann in the early 20th century, particularly his 1903 publication on states of aggregation. The concept was formalized in 1930 by Fritz , who analogized it to a second-order due to its continuous nature in thermodynamic properties.

Formal Definitions

The glass transition is formally defined as the temperature regime in which the relaxation times associated with molecular or structural rearrangements in a supercooled become comparable to the duration of the experimental , typically on the order of 100 seconds. This kinetic criterion highlights the nonequilibrium nature of the process, where the system's inability to equilibrate on accessible timescales leads to the formation of a structurally arrested, . Operationally, the glass transition temperature T_g is often specified as the isotherm at which the shear viscosity \eta attains $10^{12} Pa·s, marking the onset of solid-like behavior in the material. Equivalently, it corresponds to the point where the structural relaxation time \tau reaches approximately 100 s, or where the instantaneous transitions to a value around $10^9 , reflecting the dominance of over viscous responses. From a thermodynamic viewpoint, the glass transition manifests as a pseudo-second-order transition, featuring a step-like discontinuity in the C_p at T_g, while the first derivatives of the remain continuous; this implies a discontinuous of the , akin to classical second-order phase changes but without true singularity. The underlying configurational S_c plays a central role, as the transition arises when cooperative molecular rearrangements become too sluggish to allow the system to access its full entropy landscape, effectively freezing in a disordered and driving the out of . This entropic arrest, formalized in the Adam-Gibbs framework, links the exponential divergence of relaxation times to the diminishing S_c near T_g, emphasizing the role of structural disorder in . The concept of fictive temperature T_f provides a means to quantify this nonequilibrium state, defined as the hypothetical temperature of an supercooled that would possess the same and as the actual glass at its formation temperature. Thus, T_f serves as a structural parameter that evolves with thermal history, bridging the glassy state's frozen properties to an equivalent counterpart.

Transition Temperature

Definition and Measurement

The glass transition temperature, denoted as T_g, is operationally defined using (DSC) as the midpoint of the step change in , determined as the of the tangent to the curve at the point of inflection with the extrapolated pre- and post-transition baselines, marking the point where the material's thermodynamic response shifts from rigid to more fluid-like behavior. Alternatively, T_g can be identified from the change in the rate of volume expansion, where the coefficient of transitions from the lower value characteristic of the (\alpha_g) to the higher value of the supercooled liquid (\alpha_l). This definition aligns with operational standards for assigning T_g in amorphous materials, emphasizing the kinetic nature of the transition rather than a strict . Differential scanning calorimetry (DSC) is a primary method for measuring T_g, where the transition appears as a step-like increase in , manifesting as an endothermic baseline shift in the heat flow curve during heating. The T_g is typically determined from the of this step, often calculated as the between the extrapolated onset and end of the transition, following standardized procedures such as ASTM E1356. For precise measurement, DSC instruments are calibrated using high-purity as a reference standard, which has a well-defined melting transition at 156.5985 °C and of 28.58 J/g, ensuring accurate temperature and heat flow scaling. Dilatometry measures T_g by tracking dimensional changes with , revealing a characteristic kink in the curve where the expansion rate accelerates due to the onset of structural relaxation. Below T_g, the relative length change follows \Delta L / L_0 = \alpha_g (T - T_g), while above it, \Delta L / L_0 = \alpha_l (T - T_g), with linear fits to the glassy and regimes used to locate the defining T_g. This technique, often implemented via (TMA), is particularly sensitive for materials where volume changes dominate the transition signature, as per ISO 11359-2 guidelines. Viscometry defines T_g operationally as the temperature where the shear viscosity reaches $10^{12} Pa·s (log \eta = 12), corresponding to a relaxation time of approximately 100 seconds, measured using techniques like fiber or beam bending for high-viscosity regimes. Complementary to this, (DMA) identifies T_g from the peak in the loss tangent (tan \delta) or the maximum in the loss modulus during oscillatory testing, reflecting the temperature where mechanical relaxation broadens significantly, as standardized in ASTM E1640. These rheological methods are essential for capturing the dynamic aspects of the transition in viscoelastic materials. The measured T_g exhibits dependence on the heating or cooling rate q, with faster rates yielding higher T_g values due to the kinetic lag in structural equilibration; an empirical relation approximates this as T_g \approx T_{g0} + C \log q, where T_{g0} is the reference T_g at a standard rate (e.g., 10 K/min), and C is a material-specific constant typically ranging from 2 to 5 K per decade of rate. This rate sensitivity underscores the nonequilibrium nature of the glass transition and necessitates consistent experimental conditions for comparability across studies.

Factors Influencing Tg

The glass transition temperature (Tg) is profoundly influenced by molecular , particularly and intermolecular forces. In polymers, increased , such as from rigid aromatic groups or bulky side chains, restricts segmental mobility and raises Tg by reducing the of the system. Stronger intermolecular forces, like hydrogen bonding, further elevate Tg by enhancing cohesive interactions that hinder rearrangements during cooling. Free volume theory posits that Tg arises when the available free volume falls below a critical threshold, limiting molecular motion; materials with lower inherent free volume, due to compact packing from stiff chains or strong attractions, exhibit higher Tg. Cooling rate significantly modulates the apparent through kinetic effects on structural relaxation. Faster cooling rates increase because the system has insufficient time for complete relaxation, trapping it in a higher-energy, less equilibrated state with reduced free volume. This kinetic origin is evident in techniques like (), where shifts by approximately 3–5 K per decade change in cooling rate. Pressure dependence of Tg stems from its compression of free volume and alteration of relaxation dynamics, typically following a Clausius-Clapeyron-like relation derived from thermodynamic changes at the transition. The coefficient dTg/dP ranges from 0.1 to 0.3 K/MPa across various glass-formers, reflecting how elevated pressure slows by densifying the structure. Additives and plasticizers lower Tg by introducing excess free volume and enhancing chain mobility. Low-molecular-weight solvents or plasticizers, such as dioctyl phthalate in polymers, disrupt intermolecular forces and increase the fractional free volume, depressing Tg proportionally to their concentration. In nanoscale confinement, such as thin films or pores, often decreases by 10–50 K compared to the due to enhanced surface interactions that accelerate surface-layer and reduce overall . This effect is prominent in supported films, where free surfaces dominate, leading to a in . For mixed systems, universal mixing rules predict Tg based on component weight fractions (w_i) and pure-component Tg values. The Fox-Flory equation, applicable to miscible blends, assumes ideal volume additivity of free volume contributions: \frac{1}{T_g} = \frac{w_1}{T_{g1}} + \frac{w_2}{T_{g2}} This linear reciprocal form captures deviations from additivity in copolymers. The more general Gordon-Taylor equation accounts for differing expansivities via a fitting k (often k ≈ ρ1 Cp1 / ρ2 Cp2, where ρ is and Cp is ): T_g = \frac{w_1 T_{g1} + k w_2 T_{g2}}{w_1 + k w_2} This relation effectively models Tg in amorphous mixtures, including pharmaceuticals and composites.

Thermodynamic Aspects

Heat Capacity Changes

The glass transition is characterized by a discontinuous increase in the isobaric heat capacity C_p at the transition temperature T_g, typically on the order of \Delta C_p \approx 0.1 to $0.5 J/g·K for many amorphous materials, which signifies the onset of configurational degrees of freedom as the material shifts from a rigid glass to a more mobile supercooled liquid state. This jump arises because, below T_g, molecular rearrangements are frozen, limiting contributions to heat capacity primarily to vibrational modes, whereas above T_g, additional anharmonic and cooperative configurational excitations become active, enhancing the material's ability to absorb heat. In metallic glasses, for instance, \Delta C_p values cluster around 13-14 J/mol·K, often approximating $3R/2 (where R is the gas constant), underscoring a universal scaling tied to atomic-scale freedoms. Heat capacity in glasses can be modeled using temperature-dependent expressions that distinguish vibrational from configurational contributions. Below T_g, the of the glass follows a predominantly vibrational form, approximated as C_{p,\text{glass}} \approx a + bT, where a and b are material-specific constants reflecting vibrations akin to those in crystals, with minimal . Above T_g, the supercooled 's incorporates an additional configurational term, often modeled as C_{p,\text{liquid}} \approx a + bT + c/T, where the c/T contribution accounts for relaxational processes and structural rearrangements that inversely with . These models highlight how the configurational , frozen in the , unfreezes upon heating through T_g, leading to a step-like enhancement in response without a , distinguishing the transition from a phase change. The and changes associated with this \Delta C_p have significant thermodynamic implications for the supercooled liquid. The excess relative to the is given by \Delta H = \int_{T_g}^{T} \Delta C_p \, dT', representing the stored in configurational states as temperature rises above T_g. Consequently, the excess in the supercooled liquid accumulates as \Delta S = \int_{T_g}^{T} (\Delta C_p / T') \, dT', reflecting the increased number of accessible microstates due to molecular , which drives the material's viscous flow behavior. This excess persists in the supercooled regime, influencing and relaxation . Experimentally, changes are observed in () as a sigmoidal step in the heat flow curve over a range of about 10-20 , centered at T_g, where the baseline shifts upward due to the \Delta C_p jump. Upon reheating an aged sample, an overshoot often appears in the DSC trace near T_g, attributed to recovery as structural relaxation releases stored from the nonequilibrium glassy state. These features provide a direct probe of the transition's kinetic and thermodynamic nature. Below T_g, the vibrational in glasses exhibits an excess contribution known as the peak, typically observed in the 1-10 THz range via inelastic or , which correlates with an anomalous rise in low-temperature beyond the T^3 prediction. This peak arises from quasilocalized vibrational modes in the disordered , contributing to the vibrational and distinguishing glassy dynamics from crystalline ones, with its intensity linked to structural heterogeneity.

Kauzmann's Paradox

Kauzmann's paradox refers to a thermodynamic inconsistency that emerges when extrapolating the properties of supercooled liquids below the glass transition T_g. Proposed by Walter Kauzmann in his seminal review, the paradox highlights an apparent "entropy catastrophe" where the extrapolated configurational of the supercooled liquid would become negative at a finite , violating fundamental thermodynamic principles. The thermodynamic basis of the paradox lies in the decomposition of the liquid's entropy relative to the crystalline state. The entropy of the supercooled liquid is given by
S_\text{liquid} = S_\text{crystal} + \Delta S_\text{vib} + \Delta S_\text{conf},
where \Delta S_\text{vib} represents the difference in vibrational entropy (typically small and positive) and \Delta S_\text{conf} is the configurational entropy difference arising from the multitude of accessible molecular arrangements in the liquid. Upon cooling below the melting point, \Delta S_\text{conf} decreases more rapidly than \Delta S_\text{vib} because the heat capacity at constant pressure C_p of the liquid exceeds that of the crystal, leading to a steeper decline in entropy with decreasing temperature.
This behavior originates from the observed discontinuity in at T_g, where C_{p,\text{liquid}} > C_{p,\text{crystal}}, prompting a linear of the 's using the glassy state's properties below T_g. The predicts that the excess \Delta S = S_\text{liquid} - S_\text{crystal} reaches zero at the Kauzmann temperature T_K < T_g, implying \Delta S_\text{conf} < 0 for T < T_K, an unphysical state where the disordered would possess less than the ordered . The implications of this paradox challenge the validity of equilibrium thermodynamics for deeply supercooled liquids, as it suggests a breakdown where the liquid's disorder fails to maintain its expected entropic advantage, potentially contravening the third law of thermodynamics. In practice, glass formation circumvents the catastrophe by kinetically arresting the system at T_g, rendering the glass non-ergodic and frozen in a metastable, out-of-equilibrium configuration that does not follow the extrapolated equilibrium path. Graphically, the paradox is depicted in a plot of entropy S versus inverse temperature $1/T, where the supercooled liquid's entropy curve, characterized by a larger slope (dS/d(1/T) = -C_p/T) due to its higher C_p, intersects the shallower crystal curve at $1/T_K, visually underscoring the impending entropy crossover.

Theoretical Resolutions

One prominent theoretical resolution to the entropy paradox in glass-forming systems is provided by the Adam-Gibbs theory, which links the dramatic increase in structural relaxation time near the glass transition to the decreasing configurational entropy of the supercooled liquid. In this framework, the relaxation time \tau is expressed as \tau = \tau_0 \exp\left(\frac{A}{T \Delta S_{\text{conf}}}\right), where \tau_0 is a characteristic time scale, A is a constant related to the energy barrier for rearrangements, T is the temperature, and \Delta S_{\text{conf}} is the configurational entropy per molecule. Vitrification is interpreted as occurring when \Delta S_{\text{conf}} approaches zero at the Kauzmann temperature T_K, avoiding the unphysical negative entropy by positing that the system kinetically arrests before reaching thermodynamic instability. This entropy-driven mechanism emphasizes molecular rearrangements, with the size of rearranging regions growing as \Delta S_{\text{conf}} diminishes. The random transition (RFOT) theory extends this entropic perspective by incorporating nucleation-like mechanisms for structural rearrangements in supercooled liquids, resolving the paradox through a mosaic picture of the system. Below a dynamic crossover temperature T_d > T_g, cooperative regions of finite size form and grow via activated processes that mimic a , but randomness in the energy landscape prevents a sharp thermodynamic singularity. The configurational vanishes at T_K < T_g, but the glass transition is preempted by kinetic slowing due to the free-energy cost of creating interfaces between amorphous domains, ensuring thermodynamic consistency without crystallization. This approach highlights the role of long-range correlations and surface tension in stabilizing the amorphous state. However, simulations in certain model glass-formers, such as binary hard disk mixtures, indicate that the configurational entropy remains positive and close to the mixing entropy near the kinetic glass transition, suggesting it does not reach zero in these systems and providing a nuance to the idealized vanishing at T_K in the Kauzmann paradox. Mode-coupling theory (MCT) offers a dynamical resolution by predicting an ideal glass transition arising from the self-consistent feedback of fluctuations, which arrests long-time motion before the Kauzmann temperature is reached. In MCT, the structural relaxation function exhibits a non-ergodicity transition at a critical temperature T_c > T_g, where the long-time limit of the correlator becomes finite due to the of vertex functions in the mode-coupling equations. This dynamical arrest occurs through the coupling of collective modes, leading to caging effects that slow without invoking directly, though it complements entropic theories by placing T_c above T_K. The theory's predictions for the power-law s near T_c provide a microscopic basis for the kinetic avoidance of the crisis. Variants of entropy theory, particularly the potential energy landscape (PEL) approach, further reconcile the paradox by viewing the glass as trapped in deep minima of a rugged , circumventing low- crystalline states through kinetic barriers. In this , supercooled liquids explore an increasingly constrained set of basins as decreases, with the glass transition marking the point where is insufficient to escape higher-barrier regions, preserving positive configurational by avoiding the global minimum associated with T_K. Seminal developments emphasize how the of the PEL funnels the system into amorphous configurations, with inherent structures (local minima) dictating relaxation pathways and preventing the extrapolation from becoming negative. These theories also connect to the concept of fragility, where strong glass-formers like silica (SiO₂) exhibit T_K close to T_g due to gradual loss and modest effects, while fragile liquids show T_K farther below T_g with sharper non-Arrhenius behavior driven by larger entropic drops. This distinction underscores how theoretical resolutions adapt to material-specific landscapes, maintaining thermodynamic viability across diverse systems.

Dynamic Behavior

Time-Temperature Superposition

The time-temperature superposition (TTS) principle asserts that viscoelastic properties of glass-forming materials, such as relaxation moduli or , measured at various temperatures near the glass transition can be overlaid to form a single master curve by shifting curves horizontally along a logarithmic time (or ) axis using a temperature-dependent shift factor a_T. This approach effectively equates changes in temperature to adjustments in the observation time scale, enabling prediction of long-term behavior from short-term experiments. The shift factor a_T is quantitatively captured by the Williams-Landel-Ferry (WLF) equation, derived from free volume considerations in viscous flow: \log_{10} a_T = -\frac{C_1 (T - T_\mathrm{ref})}{C_2 + (T - T_\mathrm{ref})} where T is the measurement temperature, T_\mathrm{ref} is a reference temperature (often the glass transition temperature T_g), and C_1 and C_2 are empirical constants. For many amorphous polymers, C_1 \approx 17.44 and C_2 \approx 51.6 K when T_\mathrm{ref} = T_g, reflecting near-universal behavior across diverse systems. The equation holds reliably in the range T_g < T < T_g + 100 K, where cooperative relaxation dominates. Physically, TTS arises from the thermal activation of molecular rearrangements in the viscous regime, where increasing temperature accelerates relaxation rates in a manner analogous to extending the time scale, akin to time dilation in activated processes. This equivalence stems from the temperature dependence of configurational entropy and free volume, which govern barrier crossing for structural relaxations in supercooled liquids. The principle applies robustly to entangled polymers, where chain dynamics exhibit broad relaxation spectra, and extends to certain small-molecule glass-formers exhibiting similar cooperative dynamics above T_g. However, TTS validity diminishes at temperatures well above T_g + 100 K, as simpler, non-cooperative modes emerge and shift factors deviate from WLF predictions. Near T_g, the underlying non-Arrhenius temperature dependence—manifested in the WLF form—leads to increasingly divergent activation energies for relaxation as temperature approaches T_g from above, limiting simple Arrhenius extrapolations.

Fragility and Master Curves

The fragility of a glass-forming liquid quantifies the sensitivity of its structural relaxation time \tau (or viscosity \eta) to temperature changes near the glass transition temperature T_g. This is captured by the fragility index m, defined as m = \left. \frac{d \log_{10}(\tau)}{d(T_g/T)} \right|_{T=T_g}, where \tau_g is the relaxation time at T_g (typically \sim 100 s) and \tau_0 is a microscopic time scale (\sim 10^{-14} to $10^{-12} s). Liquids with m < 100 are classified as "strong," exhibiting Arrhenius temperature dependence (\log \tau \propto 1/T) indicative of stable, network-like structures, while those with m > 100 are "fragile," showing super-Arrhenius behavior with rapid, non-linear increases in \tau. This , introduced by Angell, highlights fundamental differences in dynamic heterogeneity and structural evolution during . For fragile liquids, the temperature dependence of \tau is often described by the Vogel-Fulcher-Tammann (VFT) equation: \log \tau = A + \frac{B}{T - T_0}, where A and B are constants, and T_0 is the Vogel temperature, typically T_0 \approx T_g - 30 to $50 K, representing a point below T_g. The parameter B/(T_g - T_0) correlates with fragility, with higher values indicating greater deviation from Arrhenius behavior and increased dynamic fragility. This form empirically captures the cooperative slowing of dynamics in fragile systems, distinguishing them from strong liquids that follow simpler activated processes. Master curves provide a predictive of relaxation across wide time scales by applying the time-temperature superposition (TTS) principle, plotting reduced quantities such as or relaxation against reduced time t / a_T, where a_T is the horizontal shift factor. These curves, spanning 10–15 orders of magnitude in time or , reveal the full of viscoelastic response from glassy to states. Experimentally, they are constructed from data obtained via techniques like (probing dipole reorientations) or (measuring shear response), with non-linear shifts in \log a_T versus T reflecting the super-Arrhenius nature, particularly pronounced in fragile . This framework exhibits universality across glass-formers: strong liquids like silica (SiO_2) have low fragility (m \approx 20), maintaining near-Arrhenius dynamics due to covalent network rigidity, while fragile liquids like display higher m \approx 53, with more pronounced effects and structural rearrangements. In contrast, highly fragile liquids or polymers can reach m > 200, underscoring the spectrum of dynamic arrest mechanisms.

Applications in Materials

Inorganic Glasses

Inorganic glasses, particularly those formed by covalent network structures such as silicates, exhibit glass transitions characterized by high temperatures and strong liquid behavior due to their rigid tetrahedral frameworks. Pure silica (SiO₂) glass represents the archetypal strong glass former, with a glass transition (T_g) of approximately 1450 K, reflecting its high network connectivity where atoms are coordinated by four oxygen atoms in SiO₄ tetrahedra. This structure leads to a fragility m ≈ 20, indicating relatively Arrhenius-like dependence near T_g, as opposed to the more non-Arrhenius behavior in fragile liquids. The addition of network modifiers like Na₂O to silica significantly alters the glass transition by disrupting the covalent network. In glasses, incorporating Na₂O breaks Si-O-Si bridges, creating non-bridging oxygens (NBOs) that increase ionic mobility and reduce structural rigidity, thereby lowering T_g to around 700 K for typical compositions such as Na₂O·3SiO₂. This modification enhances processability but can compromise thermal stability if modifier content is too high. Borate and glasses display T_g values that vary markedly with composition, often showing intermediate behavior between strong and fragile liquids. For instance, pure B₂O₃ glass has a T_g ≈ 550 K, with boron atoms in triangular BO₃ units forming a loosely connected network that exhibits strong liquid characteristics at low modifier contents but transitions toward fragility as alkali oxides are added, altering boroxol ring structures. glasses, based on PO₄ tetrahedra, typically have even lower T_g (around 500–700 K depending on P₂O₅ content and modifiers), with fragility increasing due to weaker P-O bonds and higher chain-like connectivity compared to silicates. At the structural level, the glass transition in these inorganic networks involves bending and partial breaking of Si-O or B-O bonds, enabling diffusive rearrangements without full . This process is accompanied by a jump ΔC_p ≈ 0.3 J/g·K, which is relatively small compared to polymers, underscoring the vibrational dominance in these covalent systems over configurational changes. In applications such as optical fibers, the high T_g of silica-based glasses ensures thermal stability during high-temperature drawing and operation, where low-T_g modifier components like oxides are minimized to prevent viscous flow or under service conditions up to 1000°C.

Polymers

In amorphous polymers, the glass transition temperature () marks the shift from a rigid, glassy state to a more flexible, rubbery state, driven by segmental chain motions rather than cooperative rearrangements. This behavior is exemplified by (PMMA), which exhibits a Tg of approximately 105°C and is classified as a fragile glass-former due to its pronounced non-Arrhenius relaxation dynamics near Tg. In contrast, (PC) displays a higher Tg around 145–148°C, reflecting greater chain stiffness from its aromatic backbone. Blending polymers alters Tg behavior based on miscibility. Miscible blends exhibit a single intermediate Tg, often predicted by the Gordon-Taylor equation, which accounts for compositional weighting and differences in thermal expansion. Immiscible blends, however, retain two distinct Tgs corresponding to each phase, preserving the individual transition temperatures. Compatibilizers, such as block copolymers, can broaden the transition region in partially miscible systems by enhancing interfacial interactions and reducing phase separation sharpness. Structural modifications like bulky side chains and cross-linking significantly influence Tg by restricting chain mobility. Bulky pendant groups, such as phenyl rings in derivatives, increase Tg by hindering rotational freedom along the backbone. Cross-linking further elevates Tg by forming a that limits segmental motion; for instance, cured resins often achieve Tgs exceeding 150°C, with the exact value scaling linearly with density. External factors like and residual solvents also modulate Tg. Hydrostatic raises Tg at a rate of approximately 0.2 per , as it compresses free volume and slows relaxation processes. Conversely, residual solvents act as plasticizers, depressing Tg by increasing chain mobility through hydrogen bonding or swelling effects. In industry, tailoring is essential for optimizing performance in applications like plastics and adhesives. For rigid plastics such as PMMA-based sheets, a above ensures dimensional stability, while adhesives may target values near or below ambient conditions to maintain tackiness, as in pressure-sensitive formulations. Elastomers, like , are designed with well below (around -70°C) to enable flexibility and elasticity at service temperatures.

Metallic Glasses

Metallic glasses, also known as bulk metallic glasses (BMGs) when formed in dimensions exceeding 1 mm, are typically composed of multi-component alloys involving three or more elements to enhance glass-forming ability (GFA). A representative example is the Zr-based alloy Vitreloy 1 (Vit1), with the composition Zr41.2Ti13.8Cu12.5Ni10Be22.5, which exhibits a glass transition temperature (Tg) of approximately 650 K and a crystallization temperature (Tx) of about 700 K, allowing for bulk samples larger than 1 mm via relatively low cooling rates of around 1 K/s. These alloys leverage metallic bonding and atomic packing to achieve amorphous structures, distinct from the chain-like molecules in polymeric glasses, enabling high-density, isotropic materials with superior mechanical properties. The glass transition in BMGs is often characterized by fragility parameters (m) typically in the range of 30–80, indicating intermediate fragility and a relatively sharp transition due to efficient atomic packing in multi-component systems that suppresses . This fragility arises from the cooperative dynamics in dense metallic liquids, where structural heterogeneity leads to rapid changes in relaxation times as approaches Tg. Thermal stability of BMGs is quantified by the supercooled liquid region, defined as = Tx - Tg, with values greater than 50 K identifying good glass-formers that resist during processing. This stability is guided by Inoue's empirical rules for GFA, which emphasize compositions near deep eutectics, at least three constituent elements with significant atomic size differences (>12%), and negative heats of mixing to promote atomic disorder and kinetic barriers to . For instance, Vit1 adheres to these principles, yielding ≈ 50 K and enabling practical bulk formation without rapid quenching. Below Tg, BMGs demonstrate exceptional strength, often exceeding 1 GPa, owing to the absence of dislocations and grain boundaries, but they exhibit brittle behavior under tensile loading due to localized deformation. is primarily governed by banding, where narrow zones (~10-20 nm thick) accommodate through cooperative rearrangements, leading to if multiple bands do not interact effectively. This mechanism highlights the trade-off between high elastic limits (~2%) and limited at ambient temperatures. Applications of BMGs capitalize on their unique combination of properties, including Fe-based variants for low-loss magnetic cores in transformers due to high permeability and soft . Zr- and Ti-based BMGs are employed in sporting goods, such as heads, for their high strength-to-weight ratio and impact resistance. Recent developments focus on tuning Tg and in Zr- and Mg-based BMGs for biomedical implants, offering resistance and elastic moduli closer to (~50-100 GPa) to minimize shielding.

Emerging Materials

Recent advancements in amorphous materials have highlighted the glass transition in novel systems such as two-dimensional () hybrid perovskites, where ultrafast reveals Tg values in the range of 300–400 . These materials form glasses through rapid quenching of the melt using flash (), enabling reversible glass-crystal transitions that preserve layered structures. The transition mechanism primarily involves reorientation of organic cations, which contributes to the material's dynamic behavior and low Tg, facilitating applications in and phase-change devices for and . Nanoconfinement in thin organic films significantly alters Tg due to interfacial effects, with () films exhibiting a depression of approximately 20–25 K at thicknesses around 10 compared to bulk values. This shift arises from increased interfacial free volume, which enhances chain mobility and leads to greater in thinner films, allowing cooperative rearrangements at lower temperatures. Such effects are critical for understanding the stability of ultrathin polymer coatings in applications. In chalcogenide glasses like As-Se systems, typically ranges from 400–500 K, as observed in As₂Se₃ with a value near 460 K measured by ultrafast . These materials are highly photosensitive, undergoing reversible amorphous-crystalline phase changes that enable their use in phase-change memory devices with high-speed switching and optical capabilities. The thermal stability around supports reliable operation in integrated . Bioactive glasses based on amorphous calcium phosphates exhibit Tg around 750 K (approximately 475 °C), providing a low-temperature processing window for fabricating porous scaffolds that promote bone regeneration. These materials form hydroxyapatite layers in physiological environments, enhancing osteoconductivity and integration with bone tissue for scaffold applications in tissue engineering.

Vitrification Process

Mechanics of Vitrification

Vitrification occurs when a is cooled rapidly enough to bypass , leading to a non-crystalline solid state where structural arrest happens at the glass transition temperature T_g. This process kinetically suppresses the and growth of crystalline phases by reducing the time available for atomic rearrangement during cooling, resulting in a frozen-in disordered characteristic of the supercooled . In the supercooled state approaching T_g, exhibit cooperative behavior, manifested as string-like motions or domain-like collective rearrangements among particles. These cooperative excitations involve chains of molecules moving in a correlated , with the scale \xi of these dynamically heterogeneous regions increasing as the temperature decreases toward T_g, reflecting growing spatial correlations in the slowing . The resulting glass is a non-equilibrium, metastable state relative to the crystalline form, prone to structural relaxation processes that drive it slowly toward a lower-energy configuration. Below T_g, this aging manifests through the \alpha-relaxation mechanism, where residual structural rearrangements occur over extended timescales, leading to gradual evolution of properties like volume and . To achieve without , the cooling rate must exceed a q_c, approximated by the formula q_c = \frac{\Delta T}{\tau_g} \ln \left( \frac{\tau_\mathrm{liq}}{\tau_g} \right), where \Delta T is the over which is avoided, \tau_g \approx 100 s is the structural relaxation time at T_g, and \tau_\mathrm{liq} is the relaxation time in the high-temperature regime; this ensures the cooling timescale is shorter than the incubation time for . During annealing below T_g, the recovery of excess \Delta H in the follows a stretched exponential form, characteristic of the Kohlrausch-Williams-Watts function, reflecting the distribution of relaxation times in the non-equilibrium . This nonlinear relaxation process quantifies the approach toward the metastable , with the stretching parameter capturing the heterogeneity of the aging dynamics.

Structural and Electronic Influences

The structural motifs in play a pivotal role in governing the glass transition by influencing atomic packing and stability during . In metallic glasses, icosahedral short-range order predominates, where atoms arrange in highly coordinated, polyhedral clusters that favor dense packing and resist , thereby stabilizing the amorphous state up to higher temperatures. This icosahedral motif arises from geometric frustration, preventing perfect crystalline ordering and persisting as a key feature in the supercooled liquid and glassy phases. In contrast, inorganic glasses like silica exhibit tetrahedral coordination, with atoms bonded to four oxygen atoms forming SiO₄ units connected via bridging oxygens, which maintain a continuous random . Medium-range order, encompassing correlations beyond nearest neighbors (e.g., ring distributions and chain-like arrangements in silica), endures below the glass transition temperature (T_g), contributing to the rigidity and vibrational properties of the glass while allowing subtle structural relaxation upon annealing. Electronic structure further modulates the glass transition by affecting interatomic interactions and stability. In amorphous materials, the () at the often features a pseudogap—a dip in available electronic states—that enhances glass-forming ability by reducing flexibility and promoting structural arrest. This pseudogap is particularly evident in covalent and metallic glasses, where it correlates with lower and higher resistance to . In metallic glasses involving transition elements, the degree of d-band filling influences T_g; alloys with partially filled d-bands (e.g., early transition metals like Zr) exhibit lower T_g due to more directional bonding, whereas those approaching filled d-bands (e.g., late transition metals like Cu or ) show enhanced stability and higher T_g through optimized electron delocalization and packing efficiency. The role of bond energies underscores how cohesive forces dictate T_g across material classes. Higher average bond strengths generally correlate with elevated T_g, as stronger bonds require greater to overcome during viscous flow; empirical relations link T_g linearly to mean ⟨E⟩, such as T_g ≈ 314.75(⟨E⟩ + 0.41) in select chalcogenide systems. For instance, inorganic glasses like silica benefit from robust Si-O bonds (≈452 /), yielding T_g around 1470 K (1200 °C), far exceeding polymers reliant on weaker C-C bonds (≈348 /) with T_g typically below 500 K, highlighting the transition from network to chain-like architectures. Local heterogeneity in manifests through dynamically facilitated regions (DFR), where domains propagate relaxation events cooperatively. In the DFR model, structural excitations in high-mobility sites facilitate adjacent rearrangements, leading to subdiffusive spreading of dynamics that slows overall as T_g is approached; this propagation is temperature-dependent, with stronger facilitation near T_g enhancing cooperative length scales. Such heterogeneity arises from inherent structural , linking microscopic mobility to macroscopic . A 2025 study from the elucidated the structural origins of the glass transition through liquid clustering, revealing that persistent liquid-like clusters in the supercooled state drive the boson peak—a hallmark excess vibrational mode above the prediction. These clusters, characterized by loose packing and dynamical heterogeneity, link medium-range structural motifs to the emergence of quasi-localized vibrations, providing a unified view of how atomic-scale disorder governs the transition to rigidity. This clustering mechanism persists during cooling-induced , bridging short-range order to low-frequency anomalies observed in glassy spectra.

Recent Theoretical Advances

Recent theoretical advances in understanding the glass transition have leveraged computational and analytical approaches to address longstanding questions about its fundamental mechanisms. In 2024, researchers at used simulations to investigate the glass transition in glassy systems, linking particle mobility to average squared forces and proposing a switchback model for the broad spectrum of relaxation times, providing new insights into the physics of the glassy state and breaking. These studies elucidate dynamic and structural properties during . A significant unification of phase transitions emerged in 2025 with the proposal of a "crystal-glass law," where both the melting temperature T_m and the glass transition temperature T_g are proportional to the ratio of melting enthalpy to entropy, \Delta H_m / \Delta S_m. This relation, derived from nonaffine lattice dynamics, bridges crystalline melting and amorphous vitrification by linking them through fragility and thermal expansion parameters, offering a predictive framework for transition temperatures across materials. In , a 2025 study utilized convolutional neural networks to predict T_g of polyimides from SMILES-derived molecular descriptors, achieving a of approximately 23 K. This approach enables of polymer candidates by encoding structural motifs into representations, capturing intermolecular interactions that influence without relying on experimental descriptors. Refinements to random first-order transition (RFOT) theory from 2023–2025 simulations have clarified the mosaic length scale , estimating it at approximately 2–5 near . These computational studies, incorporating dynamic heterogeneity, show that grows monotonically with decreasing , supporting RFOT predictions of cooperative regions in supercooled liquids while resolving discrepancies in length scale divergence. Despite these progresses, fundamental debates persist, particularly on whether the glass transition represents a true thermodynamic phase change or merely an apparent kinetic slowdown, as highlighted in ongoing lists of unsolved problems in physics. Recent resolutions to the Kauzmann paradox, such as those within RFOT frameworks, continue to inform these discussions but leave the intrinsic nature unresolved. A 2025 study in Nature explored thermal stresses during vitrification, finding that higher T_g values in aqueous solutions mitigate cracking risks in cryopreservation applications. By modeling stress development from differential contraction, the work shows that elevated T_g delays the onset of brittle failure, providing practical implications for biostorage while underscoring the role of transition kinetics in mechanical stability.

References

  1. [1]
    Calorimetric glass transition explained by hierarchical dynamic ... - NIH
    The glass transition refers to the nonequilibrium process by which an equilibrium liquid is transformed to a nonequilibrium disordered solid, or vice versa.
  2. [2]
    Glass Transition Thermodynamics and Kinetics - Annual Reviews
    Apr 1, 2013 · We review representative theoretical views on the unusual kinetics of liquid supercooling, which fall into two broad competing categories: thermodynamic and ...<|separator|>
  3. [3]
    The glass transition caught in the act - The Source - WashU
    Jul 17, 2017 · Glasses are weird “solid liquids” that are cooled so fast their atoms or molecules jammed before organizing themselves in the regular patterns of a crystalline ...
  4. [4]
    The glass transition and elastic models of glass-forming liquids
    Sep 29, 2006 · Basic characteristics of the liquid-glass transition are reviewed, emphasizing its universality and briefly summarizing the most popular phenomenological ...
  5. [5]
    Mode-Coupling Theory of the Glass Transition: A Primer - Frontiers
    This review outlines the key physical ingredients of MCT, its predictions, successes, and failures, as well as recent improvements of the theory. We also ...
  6. [6]
    The Glass Transition: Its Measurement and Underlying Physics | NIST
    Oct 16, 2008 · The glass transition is an important phenomenon in the practical world, where it determines the use temperature and processing temperature ...Missing: definition | Show results with:definition
  7. [7]
    Glass transition temperature from the chemical structure of ... - NIH
    Feb 14, 2020 · The glass transition temperature (Tg) is a key property that dictates the applicability of conjugated polymers.
  8. [8]
    [PDF] Glass Transformation- Range Behavior - Lehigh University
    The glass transition is a kinetic phenomena. • Thermal history dependence. • Thermal history effects on glass properties described using the. 'fictive ...
  9. [9]
    Glass Transition, Crystallization of Glass-Forming Melts, and Entropy
    The definition of a glass as advanced by Tammann and Simon has some limitations; it is strongly connected with the traditional method of producing glasses via ...Missing: 1897 | Show results with:1897
  10. [10]
    None
    Nothing is retrieved...<|separator|>
  11. [11]
    Glass transition temperature from the chemical structure of ... - Nature
    Feb 14, 2020 · The glass transition temperature (Tg) of conjugated polymers governs chain dynamics and mechanical properties. Below Tg, the storage moduli of ...
  12. [12]
    Glass Transition Temperature - NETZSCH Analyzing & Testing
    The glass transition temperature, T g , of a material characterizes the temperature range over which this glass transition occurs.
  13. [13]
    [PDF] Challenge and solution of characterizing glass transition ... - OSTI.GOV
    ... DSC with 10 K/min heating/cooling rate (Second heating scans were used for data analysis.). The dotted line and dash lines are the glass line and liquid ...
  14. [14]
    [PDF] Analysis and degradation mechanisms of enamels ... - UPCommons
    One of the most interesting texts is a dissertation wrote in 1897 ... glass transition temperature of the Modernist ... Gustav Tammann was the first academic, late ...
  15. [15]
    [PDF] The Glass Transition, Structural Relaxation and Theories of Viscosity
    The appearance of temporal heterogeneity in the relaxation of microscopic density fluctuations is a defining characteristic of the glass transition. Page 44 ...<|separator|>
  16. [16]
    Solid-that-Flows Picture of Glass-Forming Liquids - ACS Publications
    With G∞ ∼ 109 Pa, the typical ordinary liquid viscosity of 10–3 Pa·s corresponds to τα ∼ 10–12 s, which is comparable to vibration (phonon) times. On the other ...
  17. [17]
    The effect of cooling rate on the glass transition of an amorphous ...
    The glass transition has been considered either a second-order phase transition or a kinetic phenomenon. A better way to define it is a pseudo second order ...
  18. [18]
    E1356 Standard Test Method for Assignment of the Glass Transition ...
    Jan 22, 2025 · 1.1 This test method covers the assignment of the glass transition temperatures of materials using differential scanning calorimetry or ...
  19. [19]
    Dilatometry - an overview | ScienceDirect Topics
    When a glass is heated through its glass transition temperature, the structural units are free to move and the volume of the glass increases at a faster rate ...
  20. [20]
    [PDF] exploring the sensitivity of thermal analysis techniques to the glass ...
    Most DSC experiments are performed at 5 or 10ºC/minute. A higher heating rate is beneficial in detecting Tg, however, because the heat flow signal associated ...
  21. [21]
    [PDF] Certification of Standard Reference Material® 2232a
    The most popular of these, SRM 2232 Indium DSC. Calibration Standard, sold out of its existing inventory in 2017 yet continues to attract interest. To meet this ...
  22. [22]
    Phase transitions with dilatometer - Linseis
    A dilatometer can be used to determine phase transitions in materials, as a phase change is also accompanied by a change in the coefficient of thermal ...
  23. [23]
    Glass Transition Measurement - Mettler Toledo
    Learn about the glass transition, why it is of interest, its characteristics, relevant thermal analysis techniques DSC, TMA and DMA, applicable standards, ...Missing: definition viscometry
  24. [24]
    Assigning Viscosity Values in the Glass Softening Temperature Range
    Feb 14, 2023 · At around the glass transition temperature Tg, the viscosity value for glasses is around 1012.5 Pa s [3,4]. After heating, it decreases to 100 ...
  25. [25]
    Measurement of Glass Transition Temperatures by Dynamic ...
    Glass transition temperature (Tg) is measured using DMA and rheology by measuring the onset of E’/G’, the peak of E”/G”, and the peak of tan(δ).
  26. [26]
    None
    ### Summary of Empirical Relation for Tg Dependence on Heating/Cooling Rate
  27. [27]
    Dependence of a glass transition temperature on a heating rate in ...
    Jul 1, 2016 · Our result show that glass transition temperatures at low heating rates can be satisfactorily extrapolated from higher ones, with a relative ...Missing: relation | Show results with:relation
  28. [28]
    Significance of Glass Transition Temperature of Food Material in ...
    May 1, 2023 · Factors affecting glass transition temperature. The phase change from glassy to rubbery associates the value of Tg of a food material is ...
  29. [29]
    Liquids, Glasses, and the Glass Transition: A Free‐Volume Approach
    Jan 1, 1981 · Advances in Chemical Physics, Volume 48. Full Access. Liquids, Glasses, and the Glass Transition: A Free-Volume Approach ... Information. PDF icon ...
  30. [30]
    Dependence of the glass transition temperature on heating and ...
    Dependence of the glass transition temperature on heating and cooling rate ... Effect of Concentration, Cooling, and Warming Rates on Glass Transition ...
  31. [31]
    Pressure dependence of glass transition temperature in polymers
    Comparison of values for different procedures shows quite convincingly how dTg/dP, at a given heating or cooling rate, is a property not only of the polymer but.
  32. [32]
    Effect of nanoscale confinement on glass transition of polystyrene ...
    Mar 9, 2010 · One controversial and unresolved topic is the influence of characteristic size on the glass transition temperature ( T g ) for ultrathin films ...
  33. [33]
    first-principles derivation of the Gordon–Taylor equation for direct ...
    The amorphous state: first-principles derivation of the Gordon–Taylor equation for direct prediction of the glass transition temperature of mixtures; estimation ...Missing: pdf | Show results with:pdf
  34. [34]
    Heat capacity at the glass transition | Phys. Rev. B
    Jan 6, 2011 · We propose that the jump of heat capacity at T g takes place as a result of the change of the liquid's elastic, vibrational, and thermal properties.
  35. [35]
    Constant-volume heat capacity at glass transition - ScienceDirect.com
    The ΔCp in a variety of metallic glasses is almost an invariable value (13.69 J/mol/K), and is close to 3R/2 (where R is gas constant), which can be ...
  36. [36]
    Heat Capacity at the Glass Transition - IOPscience
    By identifying the configurational modes with local stresses, an explicit expression can be calculated for the heat capacity which agrees well with experiment.
  37. [37]
    [PDF] Separating the configurational and vibrational entropy contributions ...
    Apr 19, 2017 · Thermodynamically, the glass transition is indicated by a discontinuity in the specific heat as the glass relaxes to an equilibrium liquid at ...
  38. [38]
    The inquiry of liquids and glass transition by heat capacity
    Dec 20, 2012 · A continuous step in heat capacity appears across the glass transition. Clearly, the heat capacity step originates from the falling out-of ...Missing: discontinuous | Show results with:discontinuous
  39. [39]
    [PDF] Composition‐dependent glass transition temperature in mixtures
    Mar 31, 2022 · Abstract. The composition dependence of the glass transition temperature (Tg) in mix- tures remains an important unsolved problem.<|separator|>
  40. [40]
    [PDF] Overview of Glass Transition Analysis by Differential Scanning ...
    The analyst will generally define two temperature points of interest, one below and one above the glass transition step. At these two points, tangents to the ...
  41. [41]
    [PDF] Understanding Glass through Differential Scanning Calorimetry
    Aug 12, 2018 · Here we present a comprehensive review of the many applications of DSC in glass science with focus on glass transition, relaxation, ...<|separator|>
  42. [42]
    The origin of the boson peak and thermal conductivity plateau in low ...
    We have explained the microscopic origin of the excess of density of states in amorphous solids leading to the so-called boson peak in the heat capacity and a ...
  43. [43]
    Understanding the emergence of the boson peak in molecular glasses
    Jan 13, 2023 · The boson peak in the vibrational spectrum matches the excess heat capacity. As the boson peak intensifies on cooling, wide-angle x-ray ...
  44. [44]
    The Nature of the Glassy State and the Behavior of Liquids at Low ...
    The Nature of the Glassy State and the Behavior of Liquids at Low Temperatures. ... Chemical Reviews. Cite this: Chem. Rev. 1948, 43, 2, 219–256. Click to ...
  45. [45]
    Connections between some kinetic and equilibrium theories of the ...
    Apr 1, 1987 · Connections between the static density-functional theory of the glass transition and the dynamic mode-coupling theory of the glass transition are discussed.
  46. [46]
    Random first order transition theory concepts in biology and physics
    Mar 3, 2015 · In this Colloquium, we first summarize the essential ideas underlying the random first order transition (RFOT) theory of the glass transition (Article Text · Basic Notions of the RFOT for... · Glass Transition Concepts and...
  47. [47]
    The Temperature Dependence of Relaxation Mechanisms in ...
    Article July 1, 1955. The Temperature Dependence of Relaxation Mechanisms in Amorphous Polymers and Other Glass-forming Liquids. Click to copy article link ...
  48. [48]
    Time–Temperature Superposition Principle in Shearing Tests ...
    Feb 15, 2023 · The well-known principle of time–temperature superposition (TTS) is of prime interest for polymers close to their glass transition.
  49. [49]
    Correlation between Fragility and the Arrhenius Crossover ...
    Nov 10, 2016 · It is usually quantified by the kinetic fragility index m , which is defined as the slope of the Angell plot of transport coefficients in ...
  50. [50]
    A perspective on the fragility of glass-forming liquids - ScienceDirect
    Fragility focuses on a generic, striking, property of the dynamical slowdown in most glass-forming molecular liquids, polymers and inorganic materials, namely ...
  51. [51]
    Origin of the Vogel–Fulcher–Tammann law in glass-forming materials
    Often this law for polymer is written in the form log τ/τ g =C 1 T−T g T−T g +C 2 and is called the William–Landel–Ferry (WLF) relation [6], where τg is the ...Missing: original | Show results with:original
  52. [52]
    Elastic string theory understanding of the cases of glycerol and silica
    May 22, 2020 · Indeed, glycerol is a so-called intermediate glass, with fragility m = 50 – 53 [70] , against the strong character of silica, m = 19.8 [71] . ...
  53. [53]
    [PDF] A Novel View on Glass-Forming Liquids & Viscosity Model
    Nov 23, 2021 · Within this scaling description we obtain the transformed Angell plot, in which the area separating “fragile” and “strong” glass-formers emerges ...
  54. [54]
    Effect of Pressure on the Structure of SiO2 at the Glass Transition ...
    Aug 6, 2025 · The structural transformation of Silica system under compression at glass transition temperature (1475 K) investigated by molecular dynamics ...
  55. [55]
    Crystal nucleation rates in a Na2O–SiO2 glass - ScienceDirect.com
    The glass transition temperature, Tg, of the prepared glass was determined by DTA. From the DTA trace of the bulk form sample, Tg was found to be 420 ± 2°C. The ...<|control11|><|separator|>
  56. [56]
    Glass transition temperatures of binary oxides from ab initio ...
    Aug 16, 2023 · The estimated values of Tg for SiO2, GeO2, B2O3, and TeO2 are close to their measured values. The estimated values of Tg for Al2O3 and Ta2O5, ...
  57. [57]
    First-principles study on the specific heat jump in the glass transition ...
    The most important characteristic of glass transition is a jump in the specific heat ΔCp . Despite its significance, no standard theory exists to describe ...
  58. [58]
    The Role of Silica in Modern Fiber Optic Technology - FSI
    Aug 21, 2025 · Silica's low thermal expansion coefficient (≈0.55 ppm/°C) and high glass transition temperature (>1,200 °C) enable fibers to operate reliably ...
  59. [59]
    Improving glass transition temperatures of PMMA - Makevale
    Oct 22, 2024 · The main disadvantage of PMMA is its relatively low glass transition temperature (Tg) of around 105°C, making it an unsuitable material choice ...
  60. [60]
    Correlation between fragility and surface glass transition ...
    Dec 12, 2023 · The fragility of glass describes how rapidly its molecules slow down as it is cooled near its glass transition temperature.
  61. [61]
    PC: Polycarbonate - NETZSCH Analyzing & Testing
    glass transition temperatures (midpoints) are very close to each other in both heatings at 147°C (2nd heating, red) and 148°C (1st heating, blue); the step ...
  62. [62]
    Understanding Polycarbonate Glass Transition Temperature in ...
    Aug 3, 2024 · A: Polycarbonate's glass transition temperature is about 147°C. The polymeric material shifts from hard and glassy to rubbery and pliable at ...
  63. [63]
    (PDF) Prediction of glass transition temperatures: Binary blends and ...
    Aug 7, 2025 · We present a new equation for the dependence of Tg on composition in blends as well as in copolymers. We compare results obtained from earlier equations.
  64. [64]
    [PDF] 1 Glass-Transition Phenomena in Polymer Blends - Wiley-VCH
    Glass transition, or vitrification, is the freezing-in of a structural state during cooling, occurring over a narrow temperature range where molecular ...
  65. [65]
    Glass Transition Temperature (Tg) of Plastics - Definition & Values
    Aug 12, 2025 · Tg is a phenomenon of amorphous polymers. At this temperature, polymers undergo a transition from glassy to rubbery state.
  66. [66]
    8.2: Factors Influencing Tg - Engineering LibreTexts
    Sep 7, 2021 · Bulky pendant groups, such as a benzene ring, can catch on neighboring chains like a "fish hook" and restrict rotational freedom. This ...
  67. [67]
    How critical is the crosslink density in epoxies for optimizing ...
    Sep 22, 2025 · For any epoxy system, the crosslink density plays a vital role in achieving the right glass transition temperature. Not only is curing at the ...
  68. [68]
    Influence of crosslinking density on the mechanical and thermal ...
    Aug 16, 2022 · Results show that the tensile strength, Young's modulus, and glass transition temperature are linearly increased with increasing crosslinking density.
  69. [69]
    Journal of Applied Polymer Science | Wiley Online Library
    Apr 17, 2013 · McKinney and Goldstein16 showed a similar effect of pressure in a pure amorphous polymer, i.e., an increase of the glass transition temperature.
  70. [70]
    Polystyrene freeze-dried from dilute solution: Tg depression and ...
    We conclude that the residual solvent has a significant effect on the Tg depression observed for polymers freeze-dried from dilute solution; no depression or ...
  71. [71]
    Effects of Residual Solvent on Glass Transition Temperature of Poly ...
    Aug 10, 2025 · Additionally, entrapped solvents within the adhesive layer may act as plasticizing agents reducing the glass transition temperature, Tg [35] .
  72. [72]
    Glass Transition Temperature (Tg) of Polymers - Protolabs
    Glass transition temperature is the temperature at which an amorphous polymer changes from a hard/glassy state to a soft/leathery state, or vice versa.<|control11|><|separator|>
  73. [73]
    Glass Transition Temperature (Tg) in Adhesives - Permabond
    Sep 2, 2015 · Glass Transition Temperature (Tg) is the point at which a material alters state – going from a glass-like rigid solid to a more flexible, rubbery compound.Missing: tailoring elastomers
  74. [74]
    Latex Binders 101: Glass Transition Temperature
    An amorphous polymer has a glass transition temperature which is a range of temperatures across which the properties of the polymer change dramatically.
  75. [75]
    Improving glass forming ability of off-eutectic metallic glass formers ...
    From materials design viewpoint, Inoue summarizes three empirical rules that good metallic glass formers usually obey: (1) multicomponent systems consisting of ...
  76. [76]
    Review on the Research and Development of Ti-Based Bulk Metallic ...
    Johnson's group developed a good glass former Zr41.2Ti13.8Cu12.5Ni10Be22.5 (vit1) with a low critical cooling rate of ~1 K/s, which is the first commercial BMG ...
  77. [77]
    Bulk metallic glasses - ScienceDirect.com
    Vitalloy 1 (vit1), one of the most extensively studied BMG in the family, has the composition of Zr41Ti14Cu12.5Ni10Be22.5. Its temperature–time transition ...
  78. [78]
    Intrinsic correlation between fragility and bulk modulus in metallic ...
    Aug 14, 2007 · A systematic study on the available data of 26 metallic glasses shows that there is an intrinsic correlation between fragility of a liquid ...
  79. [79]
    Compositional dependence of the fragility in metallic glass forming ...
    Jun 28, 2022 · Across the considered composition space, the fragility varies smoothly (Fig. 4a). Overall, m ranges from 46 down to 16. The compositional ...
  80. [80]
    Calculations of dominant factors of glass-forming ability for metallic ...
    It is widely recognized that metallic glasses with high GFA exhibit Rc≤105 K/s, Tg/Tm≥0.6 or ΔTx≥50 K. The GFA factors have been evaluated both ...
  81. [81]
    Mechanical Properties of Metallic Glasses - MDPI
    Shear Band Formation and Non-Linear Viscosity. At room temperature metallic glasses fail by forming shear bands, where plastic strains are strongly localized.
  82. [82]
    Thickness of shear bands in metallic glasses - AIP Publishing
    Aug 15, 2006 · A review of measurements and atomistic modeling shows that shear bands in metallic glasses have a characteristic thickness of ∼10nm.
  83. [83]
    Shear bands in metallic glasses - ScienceDirect.com
    In metallic glasses, shear bands are particularly important as they play the decisive role in controlling plasticity and failure at room temperature.
  84. [84]
    (PDF) Developments and Applications of Bulk Metallic Glasses
    Metallic glass formation is achieved by passing nucleation and growth of crystalline phases when the alloy is rapidly cooled from the molten liquid. (1) In ...Abstract And Figures · References (46) · Recommended Publications
  85. [85]
    Recent advances in bulk metallic glasses for biomedical applications
    Stainless steel, vitallium (Co-Cr alloys), pure Ti and Ti alloys, pure Zr and Zr alloys are widely used as artificial hip joints, cardiovascular stents, ...
  86. [86]
    Study of Glass Formation and Crystallization Kinetics in a 2D Metal ...
    Aug 8, 2023 · The 2D hybrid halide perovskites are sorted on the basis of two key components: the inorg. layers and their modification, and the org. cation ...
  87. [87]
    Structural dynamics of melting and glass formation in a two ... - Nature
    Aug 18, 2025 · The liquid and glass formed from the two-dimensional hybrid organic-inorganic perovskite exhibit network-forming behaviour, preserving ...
  88. [88]
    Densification and depression in glass transition temperature in ...
    Oct 7, 2014 · The density of a 6 nm (0.4 gyration radius, Rg) thick film is 30% greater than that of a 150 nm (10Rg) film. A depression of 25 °C in glass ...
  89. [89]
    [PDF] Study of glass transition kinetics of As2S3 and As2Se3 by ultrafast ...
    Mar 19, 2019 · Ultrafast DSC was used to study As2S3 and As2Se3 glass transition kinetics. Fragility indices were estimated, and As2Se3 showed better ...
  90. [90]
    Novel phosphate bioglasses and bioglass-ceramics for bone ...
    Nov 15, 2024 · Bioactive glasses and glass-ceramics are biomaterials characterized by excellent bone bonding abilities. Many different bioactive glass ...
  91. [91]
    Glass Transition Temperature and Mean Bond Energy of ... - MDPI
    Good correlation is observed between the glass transition temperature and the overall mean bond energy. The dependence Tg = 314.75 (0.278 < E > + 0.41) ...Missing: organic | Show results with:organic
  92. [92]
    Under what conditions can a glass be formed?: Contemporary Physics
    Generally substances are more stable in a crystalline than in a glassy state. Therefore, to form a glass, crystallization must be bypassed.
  93. [93]
    Stringlike Cooperative Motion in a Supercooled Liquid
    Mar 16, 1998 · We observe stringlike cooperative molecular motion (“strings”) at temperatures well above the glass transition. The mean length of the strings increases upon ...Missing: domain scale ξ
  94. [94]
    Anomalous structure-property relationships in metallic glasses ...
    Apr 8, 2016 · Metallic glasses are commonly found to favor denser packing structures and icosahedral order in experiments, simulations, and theoretical models ...
  95. [95]
    Revealing the medium-range structure of glassy silica using force ...
    Dec 1, 2021 · The short-range order base structure of SiO2 is well understood, with SiO4 tetrahedral units that are interconnected via bridging-oxygen atoms ...
  96. [96]
    Structure and properties of densified silica glass: characterizing the ...
    Dec 23, 2020 · At room temperature and pressures up to 20 GPa, an essentially tetrahedral network is maintained, and structural relaxation times will lengthen ...
  97. [97]
    Correlation between glasses forming ability and density of states for ...
    Jun 15, 2020 · A pseudogap is a corresponding minimal of density of states (DOS) at the Fermi level, N (EF). Therefore, the electron density of states at ...
  98. [98]
    Electron-band theory inspired design of magnesium–precious metal ...
    Jun 13, 2017 · When alloyed with another metal, elements that have a completely- or near-filled d-band require few if any electrons to fill the band and hence ...
  99. [99]
    [PDF] SOME REMARKS ON THE GLASS-TRANSITION TEMPERATURE ...
    Jan 31, 2013 · The glass-transition temperature Tg is known to be related to cohesive energy (e.g., mean bond energy 〈E〉) in most binary and ternary ...
  100. [100]
    Dynamical Facilitation Governs the Equilibration Dynamics of Glasses
    We demonstrate that dynamical facilitation generically leads to heating driven by domain growth and cooling without domains.Missing: DFR | Show results with:DFR
  101. [101]
    Quasi-localized vibrations arise from liquid-like structures in glasses
    Jul 29, 2025 · In this study, we investigate the structural origins ... H. Tanaka. , “. Universal link between the boson peak and transverse phonons in glass. ,”.
  102. [102]
    Relationship between the boson peak and first sharp diffraction ...
    Jan 10, 2025 · Boson peak (BP) dynamics refers to the universal excitation in the terahertz region of glass. In this study, the universal dynamics of BP were ...<|control11|><|separator|>
  103. [103]
    Purdue researchers make significant advancements in ...
    Nov 4, 2024 · Purdue University Davidson School of Chemical Engineering, have published several significant studies that advance our understanding of the glass transition.Missing: machine learning landscapes ergodicity breaking
  104. [104]
    [PDF] arXiv:2503.13270v1 [cond-mat.soft] 17 Mar 2025
    Mar 17, 2025 · The paper develops relations for glass transition (Tg) and crystal melting (Tm) temperatures, showing they are proportional to liquid fragility ...
  105. [105]
    Machine Learning with Enormous “Synthetic” Data Sets: Predicting ...
    Nov 17, 2022 · The results of the study have shown that this MPNN provided three times more accurate predictions for polymer bandgap values than other MPNN ...1. Introduction · 2.2. Neural Network... · 4. Conclusions
  106. [106]
  107. [107]
    Higher glass transition temperatures reduce thermal stress cracking ...
    Jul 31, 2025 · Significant recent work has investigated transport-based methods by which to decrease the risk of thermal stress cracking during vitrification ...Missing: suppress seminal
  108. [108]
    Theoretical perspective on the glass transition and amorphous materials
    Reviews of Modern Physics article stating that the glass transition is not a thermodynamic transition but empirically defined, supporting the phenomenological nature of the viscosity definition for Tg.
  109. [109]
    Do Binary Hard Disks Exhibit an Ideal Glass Transition?
    Physical Review Letters article from 2006 discussing that configurational entropy cannot be zero for an ideal amorphous glass in binary hard disk models.
  110. [110]
    Configurational entropy of binary hard-disk glasses
    Journal of Chemical Physics article from 2007 showing that configurational entropy in binary hard-disk glasses near the kinetic glass transition is close to the mixing entropy, not zero.