Fact-checked by Grok 2 weeks ago

Amorphous solid

An amorphous solid is a rigid composed of atoms, molecules, or particles arranged in a disordered, non-repeating fashion, lacking the long-range and periodic structure found in crystalline solids. Its atomic configuration resembles that of a , with short-range order but no overall organization, resulting in a "frozen" supercooled state that maintains solidity at ambient temperatures. Amorphous solids form when a molten is cooled rapidly enough to prevent the atoms from aligning into a , a process that minimizes molecular mobility and traps the disordered structure. Unlike crystalline solids, which exhibit anisotropic properties due to their ordered lattices, amorphous solids are isotropic, displaying uniform physical characteristics in . They lack sharp points, instead softening gradually over a range, and produce irregular fractures without defined planes when broken. patterns of amorphous solids are diffuse and poorly defined, reflecting their absence of long-range order, in contrast to the sharp peaks seen in . These materials can span various classes, including ceramics, polymers, metals, and semiconductors, and often demonstrate enhanced solubility, reactivity, or mechanical resilience compared to their crystalline counterparts in specific contexts. Common examples of amorphous solids include silica glass (such as window glass or , SiO₂), which forms the basis for optical fibers and laboratory ware due to its transparency and chemical inertness. Many polymers, like , adopt amorphous structures and are widely used in and insulation for their flexibility and lightweight nature. Other instances encompass natural volcanic glass like , a silica-rich (SiO₂) material, and synthetic amorphous metals, such as those in Liquidmetal® alloys (e.g., zirconium-beryllium-titanium-copper-nickel compositions), prized for their high hardness and elasticity. Even everyday items like exemplify amorphous solids in food applications. Amorphous solids play a critical role in modern technology owing to their unique properties, enabling applications across diverse fields. In electronics, amorphous semiconductors like hydrogenated amorphous silicon (a-Si:H) are essential for thin-film solar cells and large-area displays due to their cost-effective deposition and optoelectronic performance. Amorphous metals find use in high-wear coatings for industrial equipment, such as refinery components, and in precision tools like medical scalpels, where their low friction and durability outperform traditional metals. In pharmaceuticals, amorphous forms enhance drug solubility and bioavailability, accelerating dissolution rates for better therapeutic efficacy. Additionally, their optical clarity and resistance to crystallization make them ideal for lenses, while in materials science, they exhibit intriguing phenomena like low-energy excitations absent in crystals, influencing research into elasticity and lattice dynamics.

Definition and Etymology

Definition

An amorphous solid is a non-crystalline solid characterized by a disordered or molecular that lacks long-range order and , while maintaining short-range order where neighboring atoms adopt locally preferred coordination geometries similar to those in crystalline counterparts. This absence of periodicity means amorphous solids do not exhibit the sharp diffraction patterns typical of crystals when analyzed by scattering, instead producing broad, diffuse halos. In contrast to crystalline solids, which feature a periodic lattice arrangement leading to well-defined geometric shapes, sharp melting points, and anisotropic properties, amorphous solids display irregular external forms, gradual softening over a temperature range, and isotropic behavior due to their structural randomness. Unlike liquids, which flow under applied shear stress because their molecules can freely rearrange, amorphous solids resist such deformation and maintain a fixed shape, qualifying them as true solids despite their liquid-like atomic disorder. Representative examples of amorphous solids include silica glass used in windows, polymers like in plastics, metallic glasses such as iron-boron alloys valued for their magnetic properties, and formed by rapid quenching of . The ideal glass represents a theoretical non-equilibrium state in which a supercooled is kinetically arrested below its freezing point, preserving a frozen snapshot of the liquid's disordered configuration while exhibiting solid-like mechanical rigidity.

Etymology

The term "amorphous" derives from the a- ("without") and morphḗ ("form" or "shape"), literally meaning "shapeless" or "without form," and entered scientific English via Modern Latin amorphus around 1731. This linguistic root reflected early perceptions of materials lacking the ordered, geometric structure of crystals, distinguishing them from well-formed minerals. Although the specific term "amorphous" emerged in the , provided one of the earliest documented observations of non-crystalline materials in his 1665 work , where he examined the microscopic texture of and noted its uniform, non-faceted appearance under magnification, contrasting it with crystalline substances like salts. In , "amorphous" was applied to naturally occurring substances that lacked the external crystal faces typical of minerals, marking the term's initial adoption to categorize shapeless geological materials. By the , the concept was formalized in through studies of physical properties, with Hermann Kopp employing "amorphous" to describe non-crystalline forms of substances such as carbon, whose heat capacities deviated from those of their crystalline counterparts, thus broadening the term beyond mere visual description to include thermal and structural behaviors. In the , amid rapid advances in glass science and materials research, "amorphous solid" evolved to encompass a wider array of synthetic and processed materials, including polymers and metallic glasses, reflecting a shift from mineralogical origins to interdisciplinary applications in physics and .

History

Early Observations

Ancient civilizations utilized amorphous solids, particularly , long before their structural nature was understood. In , faience—a quartz-based material coated with a glassy, vitreous —was produced as early as 4000 BCE, serving decorative and functional purposes such as beads and amulets. By around 1500 BCE, true objects, including vessels and ornaments, emerged in and , crafted through melting silica with fluxes to form a non-crystalline material valued for its transparency and moldability, though artisans lacked knowledge of its atomic disorder. The development of in the enabled initial empirical observations of irregular structures in organic materials. , in his seminal 1665 publication , examined gums, resins, and similar tenacious substances, describing how they formed elongated, cohesive filaments without the regular facets or geometric patterns typical of , attributing this to the "congruity of parts" in their fluid-like yet solid states. These findings highlighted the absence of ordered arrangements in such natural products, contrasting them with crystalline minerals observed contemporaneously. By the 19th century, chemists systematically identified substances that resisted crystallization, marking a shift toward recognizing amorphous solids as a distinct class. Jöns Jacob Berzelius isolated amorphous silicon in 1824 through the reaction of silicon tetrafluoride with potassium, noting its powdery, non-faceted form in contrast to crystalline silicon prepared later. Subsequently, Thomas Graham introduced the term "colloid" in 1861 to describe gelatinous materials like gels, rubber, and glue, which diffused slowly and failed to form crystals upon solidification, exhibiting jelly-like consistency due to their dispersed, non-crystalline organization. The term "amorphous," derived from Greek roots meaning "without form," was applied to such substances to denote their lack of crystalline structure. Early 20th-century experiments advanced understanding through controlled formation of amorphous states. Gustav Tammann conducted studies on undercooling liquids below their freezing points, demonstrating in works from the to that sufficient prevented and , yielding stable glassy solids from melts of metals, salts, and organics; for instance, he quantified rates as a function of undercooling depth, establishing kinetic barriers to . These observations laid groundwork for viewing as undercooled liquids frozen in disordered configurations.

Modern Developments

In the early , significant theoretical progress was made in modeling the atomic structure of amorphous solids, particularly silicate glasses. In 1932, William H. Zachariasen proposed the continuous random (CRN) model, which describes the structure of glasses as a continuous, randomly connected of polyhedra similar to those in their crystalline counterparts, but lacking long-range translational and orientational . This model emphasized short-range while allowing for distortions in bond angles and lengths, providing a foundational for understanding glass formation based on empirical rules such as limiting oxygen coordination to no more than two cations. Zachariasen's CRN paradigm was experimentally validated shortly thereafter through studies by B.E. Warren in 1934, confirming the absence of sharp Bragg peaks and the presence of diffuse scattering indicative of random atomic arrangements. A major experimental breakthrough occurred in the late 1950s with the discovery of metallic glasses, expanding the scope of amorphous solids beyond traditional oxide glasses. In 1960, Pol Duwez and colleagues at the successfully produced the first metallic glass by rapidly quenching a molten Au75Si25 alloy at cooling rates exceeding 106 K/s using a splat-quenching technique, preventing and yielding a non-crystalline metallic structure. This innovation demonstrated that metallic systems could form amorphous phases under extreme non-equilibrium conditions, revealing unique properties such as high strength, elasticity, and corrosion resistance due to the absence of grain boundaries. The rapid quenching method paved the way for synthesizing a variety of amorphous metal alloys, influencing applications in magnetic materials and biomedical devices. During the and , experimental investigations uncovered universal vibrational and thermal anomalies in amorphous solids, distinguishing them from crystalline counterparts. The boson peak, a prominent excess in the at frequencies around 1-3 THz, was first identified in low-frequency spectra of glasses, such as vitreous silica, highlighting deviations from Debye's model due to disorder-induced quasi-localized modes. Concurrently, low-temperature studies (below 1 ) revealed universal properties including a linear specific heat term proportional to (Cp ∝ T) and a quadratic dependence of thermal conductivity (κ ∝ T2), attributed to an ensemble of two-level tunneling systems (TLS) arising from atomic tunneling between metastable configurations in the disordered structure. These findings, consolidated in the soft-potential or tunneling model proposed in 1972, underscored the role of structural heterogeneity in governing low-energy excitations across diverse amorphous materials like polymers, oxides, and metals. From the 1970s onward, advances in amorphous semiconductors and computational modeling have deepened insights into their electronic and structural behaviors. Hydrogenated amorphous silicon (a-Si:H), first developed in the 1970s through and (PECVD), emerged as a key material for thin-film transistors and . Research in the 1990s focused on defect passivation and alloying (e.g., with ) to tune bandgaps and mitigate light-induced via the Staebler-Wronski . Computational simulations, leveraging with empirical potentials like the BKS model for silica (introduced in 1990), enabled large-scale modeling of glass formation and relaxation processes, revealing medium-range order and topological constraints absent in earlier models. In the 2020s, has revolutionized structure prediction, with generative frameworks automating the design of amorphous compositions and predicting properties like glass-forming ability from atomic datasets, achieving high accuracy in simulating complex oxides and alloys without exhaustive simulations. Recent 2025 studies have further advanced the field, exploring two-dimensional amorphous materials approaching the single-layer limit and using to reveal hidden structural rules, such as soft regions embedded in medium-range order. These AI-driven approaches, often integrated with calculations, facilitate rapid exploration of metastable phases, accelerating discoveries in and .

Structure and Formation

Atomic and Molecular Arrangement

Amorphous solids exhibit short-range order in their atomic and molecular arrangements, characterized by well-defined local bond lengths and angles that closely resemble those in their crystalline counterparts, but they lack the long-range translational periodicity that defines . This local structural similarity arises from the tendency of atoms to adopt energetically favorable nearest-neighbor configurations during the solidification process, while the absence of a periodic results from rapid cooling that traps the material in a metastable, disordered . A foundational model for understanding this arrangement in oxide glasses is the continuous random network proposed by Zachariasen, which posits that the structure consists of corner-sharing polyhedra, such as tetrahedra in silicate glasses, forming an extended without periodic repetition or energy wells that would favor . In this model, the oxygen atoms bridge the network-forming cations, ensuring that no oxygen is bonded to more than two cations and that the polyhedra share corners rather than edges or faces to maintain stability, while avoiding small rings (fewer than six members) that would impose undue strain. This framework explains the isotropic nature of glasses and their ability to form over a range of compositions without sharp melting points. The (RDF), which quantifies the probability of finding atoms at a given from a reference atom, captures this structural in amorphous solids by displaying sharp at short distances corresponding to first- and second-neighbor shells, followed by oscillations that gradually dampen without the persistent periodicity seen in . These decaying oscillations reflect the preservation of coordination environments beyond which the structural correlations fade, providing a quantitative measure of the disorder's extent. In metallic glasses, for instance, the RDF reveals subtle shifts in peak positions during relaxation processes that align with changes in compositional short-range ordering. Topological constraints further elucidate the mechanical implications of this atomic arrangement through the Phillips-Thorpe theory of rigidity percolation, which treats the network as a of bonds and angles imposing constraints on atomic motion. In three dimensions, the network achieves optimal rigidity when the average reaches approximately 2.4, marking the where floppy modes disappear and the structure transitions from underconstrained to stressed-rigid, influencing properties like elasticity without relying on long-range order. This mean-field approach highlights how local connectivity determines global rigidity in covalent glasses such as chalcogenides. In amorphous polymers, the concept of voids and free accounts for the interstitial spaces arising from inefficient chain packing, where free volume represents the unoccupied regions that enable segmental motion and . These voids, distributed non-uniformly, contribute to the material's and are quantified as the excess volume beyond a hypothetical close-packed state, playing a key role in phenomena like the by providing the space necessary for cooperative rearrangements. The free volume fraction typically ranges from 2.5% to 5% in polymers like , decreasing with cooling and influencing transport properties.

Formation Processes

Amorphous solids, also known as , form through that suppress by exploiting kinetic barriers or thermodynamic instabilities, preventing the material from achieving long-range atomic order during solidification or phase transitions. These methods rely on rapid structural arrest, where the system is driven into a metastable amorphous far from . The choice of depends on the material's and desired , with kinetic factors like cooling rates dominating in melt-based techniques and thermodynamic driving forces playing key roles in solid-state routes. One primary method is rapid quenching from the melt, which achieves extremely high cooling rates to bypass and growth of crystalline phases. In melt-spinning, a molten is ejected onto a rotating chilled , attaining cooling rates exceeding 10^6 K/s, as demonstrated in the production of metallic like Fe-based alloys. This kinetic suppression of is thermodynamically favored in multi-component systems with deep eutectics, where the liquid increases rapidly near the temperature. Vapor deposition techniques, such as and thermal , produce amorphous thin films by condensing vaporized atoms or molecules onto a under vacuum conditions that limit atomic mobility. involves bombarding a target with ions to eject material, forming dense amorphous layers in materials like a-Si or chalcogenide films, while heats the source to generate a vapor for deposition. These processes enable control over film thickness and composition at rates from nm/s to μm/min, with amorphicity arising from the low temperatures that hinder diffusive rearrangement. Solid-state amorphization transforms crystalline precursors into amorphous phases without melting, driven by accumulated defects or interdiffusion that destabilize the lattice. introduces high-energy ions into a solid, creating collision cascades that disrupt crystallinity, as seen in where doses above 10^14 ions/cm² yield fully amorphous layers up to 100 nm thick. Mechanical alloying via ball milling repeatedly deforms and fractures powder particles, promoting amorphization in alloys like Zr-Al through shear-induced mixing and stored energy exceeding 10 kJ/mol. These methods highlight thermodynamic pathways where defect accumulation lowers the barrier to the amorphous state. The sol-gel process synthesizes amorphous ceramics from solution precursors, involving and of metal alkoxides to form a colloidal sol that gels into an oxide network, followed by drying and . This wet-chemical route produces amorphous silica or gels at low temperatures (<100°C), with porosity tunable via pH and aging, offering advantages over melt methods for complex oxides. Pressure-induced amorphization, conversely, compresses crystalline solids beyond their stability limit, as in ice Ih transforming to high-density amorphous ice (HDA) at pressures above 1 GPa and temperatures below 150 K, where mechanical instability drives the polyamorphic transition. The ease of forming amorphous solids via these processes correlates with the fragility index of the parent liquid, a measure of how steeply the viscosity rises near the glass transition temperature (T_g). Strong liquids, like SiO_2 with fragility m ≈ 20, exhibit Arrhenius-like behavior and high glass-forming ability due to stable network structures resisting structural relaxation, requiring modest cooling rates (~1 K/s). Fragile liquids, such as o-terphenyl with m > 100, show non-Arrhenius divergence and poorer glass-forming ability, necessitating faster (>10^4 K/s) to avoid , as fragility reflects the liquid's sensitivity to temperature changes in configurational .

Fundamental Properties

Thermal Properties and Glass Transition

Amorphous solids exhibit distinct thermal behaviors that differ markedly from crystalline materials, particularly around the glass transition temperature T_g, where the material undergoes a kinetic of structural relaxation. This transition marks the point at which the supercooled liquid's reaches approximately $10^{12} Pa·s, effectively freezing atomic or molecular motions on experimental timescales, typically occurring in the range of $10^2 to $10^3 K depending on the material composition, such as lower values for organic polymers and higher for inorganic glasses like silica. Unlike a true thermodynamic , the is rate-dependent, with T_g shifting to higher values under faster cooling rates due to incomplete relaxation. A hallmark of the glass transition is the discontinuous jump in \Delta C_p at T_g, reflecting the onset of configurational contributions to in the supercooled liquid state above T_g. This jump typically amounts to approximately 0.5 to 1 times the R per atom, providing a measure of the unlocked during the transition, and is linked to the Prigogine-Defay relation, which assesses the consistency of the transition through ratios involving \Delta C_p, , and changes. Below T_g, the glass enters a non-equilibrium state, leading to relaxation and physical aging, where stored excess is gradually released over time as the evolves toward a more stable configuration, often resulting in increased density and reduced free volume. This aging process is thermally activated and can significantly alter properties like mechanical strength, with relaxation times following Vogel-Fulcher-Tammann behavior. The dynamics near T_g are further characterized by the fragility parameter m, defined as m = \left. \frac{d \log \tau}{d (T_g / T)} \right|_{T = T_g}, where \tau is the structural relaxation time; this classifies glass-formers as "" (low m \approx 20-40, Arrhenius-like behavior, e.g., SiO_2) or "fragile" (high m > 100, strongly non-Arrhenius, e.g., many liquids), influencing the sharpness of the and ease of glass formation. Underlying these kinetic aspects is the Kauzmann , which arises from extrapolating the supercooled liquid's below T_g to a hypothetical Kauzmann temperature T_K, where the liquid's configurational would equal that of the , potentially leading to an entropy crisis; this is averted in practice by the kinetic freezing at T_g > T_K, preventing the unphysical negative excess. At low temperatures well below T_g, amorphous solids show deviations from the in specific heat, attributed to two-level systems.

Mechanical and Elastic Properties

Amorphous solids exhibit distinct elastic properties compared to their crystalline counterparts, primarily due to their disordered atomic structure. The , which measures the negative ratio of transverse to axial under uniaxial , typically ranges from 0.276 to 0.409 in metallic , often higher than in many crystalline metals (around 0.25–0.35) because the lack of long-range order allows greater lateral expansion under compression. This elevated value arises from the isotropic nature of the disorder, enabling more uniform deformation responses. The G and K characterize the resistance to and volumetric deformation, respectively, and are linked to the network's average \langle z \rangle through rigidity counting. In amorphous solids, rigidity emerges when \langle z \rangle exceeds the Maxwell threshold of $2d (where d is the dimensionality, typically 3 for materials), with G vanishing at this isostatic point while K remains finite due to its dependence on radial constraints. For metallic , G is approximately 30% lower than in corresponding , reflecting softer modes from structural , whereas K is only about 6% reduced, highlighting the relative stability of compressive responses. Plasticity in amorphous solids, particularly metallic , occurs through localized zones (STZs), which are cooperative clusters of 200–700 atoms that undergo irreversible shearing under applied . These STZs nucleate and propagate as the primary mechanism of plastic flow below the , leading to banding and enabling high strength but limited in many cases. The size and activation of STZs, typically 2.5–6.6 nm³, depend on sensitivity and correlate with overall deformability. The transition between and in amorphous solids is governed by the role of free volume in the yielding process, where higher free volume facilitates STZ activation and distributed , promoting , while lower free volume concentrates deformation into brittle shear bands. In metallic glasses, increased free volume—often tuned by processing conditions like cooling rate—enhances the critical for yielding by allowing rearrangements without catastrophic . This contrasts with brittle behavior in low-free-volume states, where shear bands propagate rapidly, leading to sudden failure. Under sustained or cyclic loading, amorphous solids display and , where creep involves time-dependent deformation via gradual STZ accumulation under constant , resulting in sublinear growth. , induced by cyclic es, promotes crack initiation along shear bands and surface modifications, often reducing endurance limits compared to crystalline alloys due to localized damage accumulation. Near the temperature, both shear and bulk moduli soften, influencing these long-term responses, though ambient-temperature dominate typical applications.

Low-Temperature Universal Behavior

At low temperatures, typically below 1 K, amorphous solids exhibit universal thermal and acoustic anomalies that distinguish them from crystalline counterparts, primarily attributed to the presence of tunneling two-level systems (TLS). These TLS arise from atoms or groups of atoms that can between two nearly degenerate positions in the disordered , leading to weakly bound excitations with energies on the order of thermal energies at cryogenic conditions. This behavior is observed across a wide range of insulating , such as silica and polymers, independent of their , highlighting the intrinsic role of structural . The specific heat of amorphous solids at these temperatures includes a linear term C = \gamma T, where \gamma is the arising from TLS excitations, typically in the range of $10^{-4} to $10^{-3} J/mol K². This contrasts with the cubic T^3 dependence expected from contributions in Debye theory for crystals, as the TLS provide a constant at low energies. Experimental measurements on materials like vitreous silica confirm this linear contribution dominates below 0.5 K, with \gamma values around 0.5 mJ/mol K² for many . The tunneling model for TLS posits a of these systems in energy splitting \varepsilon and tunneling parameter \Gamma, given by P(\varepsilon, \Gamma) = P_0, where P_0 is a material-independent constant on the order of 1 per eV per cm³. Here, \varepsilon = \sqrt{\Delta^2 + \Gamma^2}, with \Delta as the asymmetry between the two wells. This of uniformity in the parameter space explains the saturation of low-energy excitations and their weak coupling to phonons via deformation potentials, enabling the linear specific heat and other universals. The model, originally proposed for ionic tunneling, has been validated through fits to thermal data across diverse amorphous materials. Thermal conductivity \kappa in amorphous solids follows a T^2 dependence at very low temperatures (below ~1 K), resulting from phonon scattering by TLS. Long-wavelength phonons resonant with TLS energy splittings cause resonant scattering, while one-phonon relaxation processes contribute, leading to this quadratic regime before a plateau emerges around 5-10 K. Measurements on fused silica and other glasses show \kappa / T^2 values spanning nearly three orders of magnitude but universally scaling as T^2 in the quantum regime, underscoring the TLS dominance. At slightly higher temperatures (~10-50 K), the vibrational in amorphous solids reveals a boson peak, an excess over the \omega^2 prediction, centered at frequencies \omega_b \approx 1-3 THz. This feature, observed via Raman or in materials like silica glass, corresponds to quasi-localized modes arising from disorder-induced anharmonicities, contributing to enhanced low-frequency excitations. The peak height and position vary modestly with glass type but remain a hallmark of amorphous vibrational spectra. Acoustic properties further illustrate these universals, with showing a linear T dependence at millikelvin temperatures due to TLS-phonon interactions, transitioning to a in internal around 5-20 K. This peak, prominent in ultrasonic measurements on glasses like , reflects thermal relaxation of TLS resonant with frequencies, with the friction Q^{-1} reaching ~10^{-3} near 10 K before decreasing. Such behavior is nearly , though peak positions shift slightly with frequency and material.

Characterization Techniques

Diffraction and Scattering Methods

and methods are essential for characterizing the structure of amorphous solids, which lack long-range translational and thus do not produce sharp Bragg peaks characteristic of crystalline materials. Instead, these techniques reveal diffuse patterns that provide information on short- and medium-range through the analysis of broad halos and diffuse features. By employing X-rays, neutrons, or electrons as probes, researchers can extract structural metrics such as the (RDF) or pair distribution function (PDF), which describe the probability of finding atoms at specific interatomic distances. These methods average over relatively large volumes, typically on the order of nanometers cubed, offering ensemble-averaged insights into the disordered arrangements. X-ray diffraction (XRD) is a primary technique for studying amorphous solids, where the diffraction patterns exhibit broad, symmetric halos rather than discrete peaks, reflecting the absence of periodic planes. The position and width of these halos correspond to the average interatomic distances and structural disorder, respectively. To quantify local structure, the total scattering data from XRD can be Fourier-transformed to obtain the pair distribution function (PDF), which provides a real-space representation of atomic pair correlations up to several nanometers. This approach has been instrumental in analyzing materials like amorphous silica and metallic glasses, enabling the identification of coordination numbers and bonding environments. Synchrotron-based XRD enhances resolution due to higher flux and energy, allowing for finer details in the PDF. Neutron diffraction complements XRD by offering sensitivity to light elements and isotopic variations, which is crucial for amorphous solids containing , oxygen, or other low-atomic-number atoms where X-ray scattering lengths are similar. The technique's isotope sensitivity arises from the distinct neutron scattering lengths of isotopes like and , allowing selective probing of specific atomic species in multicomponent systems. Total scattering neutron diffraction captures both Bragg-like (though diffuse) and diffuse components, enabling the extraction of the RDF through of the structure factor over a wide Q-range. This has been applied to study amorphous alloys and polymers, revealing chemical short-range order that is obscured in X-ray data. Facilities like spallation sources provide high-intensity s for such measurements. Electron diffraction, particularly in transmission electron microscopy (TEM), allows for the investigation of local order in amorphous solids at the nanoscale by selecting small areas, typically 10-100 in diameter, through apertures. Selected area electron diffraction (SAED) patterns from amorphous regions show ring-like diffuse , indicative of isotropic short-range order without crystalline domains. The high spatial resolution of electron beams enables mapping of structural heterogeneity in thin samples, such as amorphous thin films or nanoparticles. Analysis of these patterns via radial integration yields intensity profiles similar to , from which reduced PDFs can be derived to assess local atomic arrangements. This method is particularly useful for confirming amorphicity in regions too small for bulk diffraction techniques. Wide-angle X-ray scattering (WAXS) extends the utility of X-ray methods by focusing on higher scattering angles (typically 5° to 50° 2θ), which probe medium-range order (up to 1-2 nm) in amorphous solids beyond the short-range correlations captured by small-angle scattering. WAXS patterns feature broad peaks that reflect correlated atomic arrangements over multiple coordination shells, such as in chalcogenide glasses or metallic amorphous alloys. By combining WAXS with PDF analysis, researchers can distinguish polyamorphic phases with differing medium-range packing densities. This technique is often performed in situ during processing, like rapid quenching, to track structural evolution. Despite their power, and methods have inherent limitations when applied to amorphous solids. The absence of sharp Bragg peaks precludes direct indexing of parameters, requiring indirect modeling to interpret diffuse . Moreover, these techniques inherently average the structure over the probed , which for or setups can span micrometers, potentially masking nanoscale heterogeneities. High-resolution variants like aberration-corrected TEM mitigate this to some extent but are limited to thin specimens. Complementary spectroscopic methods can provide additional site-specific information to overcome these averaging effects.

Spectroscopic and Absorption Techniques

Spectroscopic and absorption techniques are essential for probing the local electronic, vibrational, and chemical environments in amorphous solids, where long-range order is absent. These methods provide insights into short-range atomic arrangements and bonding without relying on periodic structures, complementing global structural analyses like in one sentence. Key techniques include (XAFS), Raman and infrared (IR) , (NMR), and (XPS), each targeting specific aspects of disorder and local heterogeneity. XAFS encompasses extended X-ray absorption (EXAFS) and X-ray absorption near-edge structure (XANES), offering element-specific information on the local coordination environment around absorbing atoms in amorphous materials such as oxide glasses. EXAFS analyzes the oscillatory beyond the , yielding precise interatomic distances with resolutions down to approximately 0.01 and coordination numbers within ±10%, as demonstrated in studies of and glasses containing transition metals like and . For instance, in amorphous Fe-doped s, EXAFS has measured Fe-O bond lengths of 1.87–1.92 , revealing variations due to local distortions absent in crystalline counterparts. XANES, focusing on the near-edge region, elucidates and oxidation states through features like pre-edge peaks and edge shifts; in Ti-containing glasses, pre-edge intensities indicate average coordination numbers of 5.4–5.8, reflecting tetrahedral-to-octahedral transitions under or compositional changes. Raman and IR spectroscopy interrogate vibrational modes, which are broadened by structural disorder in amorphous solids, providing signatures of short-range bonding and phonon density of states. In crystalline materials, selection rules limit active modes to specific symmetries, producing sharp peaks, whereas amorphous solids exhibit contributions from the full phonon spectrum, resulting in broad bands with widths of several hundred cm⁻¹ due to distributions in bond lengths and angles. For example, in , Raman spectra show a transverse optical mode broadened to below 550 cm⁻¹, contrasting the narrow 520 cm⁻¹ peak in , while IR complements by detecting dipole-active modes influenced by local asymmetry. These techniques have been applied to inorganic glasses and epitaxial films, such as amorphous GeTe, where combined far-IR and Raman reveal shifts in low-frequency modes indicative of disorder-induced softening. NMR spectroscopy captures distributions that highlight site-to-site heterogeneity in amorphous solids, arising from varied local magnetic environments. In solid-state NMR, broadened lineshapes—typically 2–6 ppm for ¹³C and 0.6–1.8 ppm for ¹H—reflect conformational and intermolecular variations, as seen in amorphous pharmaceuticals like AZD4625, where enhances resolution to map bonding networks. By comparing experimental shifts to predicted distributions from molecular simulations, NMR identifies dominant local motifs, such as clustered versus isolated molecular sites, enabling atomic-level structure determination in non-crystalline drugs. XPS determines surface composition and oxidation states in amorphous solids by measuring photoelectron binding energies from the top few nanometers. In amorphous alloys like Cu₅₀Ti₅₀, reveals enriched Ti oxides at the surface with binding energies indicating Ti⁴⁺ states, differing from bulk compositions due to preferential oxidation during preparation. Similarly, for Fe₇₀Cr₁₀P₁₃C₇ amorphous ribbons, quantifies passive layers incorporating P, supporting models of resistance through semiquantitative elemental ratios and analysis. These techniques excel at elucidating short-range order in amorphous solids, where their local probes—nanometer-scale coherence lengths and element selectivity—overcome limitations of methods requiring periodicity, as in Raman's sensitivity to bond-level disorder in electroceramics.

Microscopy and Imaging Methods

Microscopy and imaging methods play a crucial role in visualizing the nanoscale heterogeneity inherent to amorphous solids, where the lack of long-range order makes traditional crystallographic techniques insufficient. These approaches provide direct spatial information on atomic arrangements, surface features, and structural variations at resolutions down to the atomic scale, revealing medium-range order and defects that influence material properties. Unlike averaged signals from diffraction or spectroscopy, imaging techniques capture local variations, enabling the study of non-uniformity in materials such as glasses, polymers, and metallic alloys. Atomic electron tomography (AET) has emerged as a powerful technique for reconstructing three-dimensional atomic positions in thin samples of amorphous solids, overcoming the challenges posed by their disordered structures. By acquiring a series of two-dimensional projections from multiple tilt angles using aberration-corrected scanning transmission electron microscopy (STEM), AET employs iterative algorithms to determine the 3D coordinates of individual atoms, achieving localization precision better than 1 Å in materials like amorphous tantalum oxide. This method has been applied to determine the full atomic structure of an amorphous solid for the first time, demonstrating liquid-like packing with short-range order similar to the parent liquid. Recent advancements, such as ptychographic AET, further enhance resolution by incorporating phase information from overlapping probes, allowing visualization of sub-angstrom features in beam-sensitive amorphous samples. Fluctuation electron microscopy (FEM) quantifies medium-range order in amorphous solids by measuring the variance in the scattering intensity from nanometer-scale volumes, providing insights into structural correlations beyond short-range atomic packing. In this , dark-field images are acquired at various probe positions, and the normalized variance of the scattered intensity serves as a measure of paracrystalline-like order, with characteristic length scales typically 0.5–2 nm detectable in materials such as and . FEM has revealed consistent medium-range order in hydrogenated films regardless of deposition method, correlating with enhanced stability against the Staebler-Wronski effect. The method's sensitivity to nanoscale heterogeneity makes it particularly useful for distinguishing subtle structural differences in non-crystalline semiconductors and oxides. Scanning tunneling microscopy (STM) offers atomic-resolution imaging of surface topography and defects in amorphous solids, leveraging quantum tunneling currents to map electronic density variations at the vacuum-solid . For instance, in hydrogenated surfaces, STM reveals local protrusions and depressions associated with dangling bonds and voids, providing direct correlation between surface morphology and bulk defect formation during growth. This technique has been used to observe defect evolution in films, where irradiation-induced migrations lead to observable rearrangements and topographic changes. STM's ability to operate under conditions minimizes contamination, enabling precise studies of surface relaxations in metallic glasses and chalcogenide alloys. Cryogenic transmission electron microscopy (cryo-TEM) is essential for imaging beam-sensitive amorphous solids like and biomolecules, preserving their native hydrated or vitrified states through rapid freezing to form . In polymer blends, cryo-TEM combined with 4D-STEM visualizes and nanoscale domains in amorphous-crystalline mixtures, such as polystyrene-block-polybutadiene, with resolutions sufficient to track morphological evolution during processing. For biomolecules embedded in amorphous matrices, cryo-TEM captures structural details of protein aggregates or vesicles without dehydration artifacts, as seen in studies of vitrified solutions revealing oligomeric heterogeneity. This approach extends to functional materials, where low-temperature operation reduces and maintains structural integrity during observation. Despite these advances, microscopy of amorphous solids faces significant challenges, particularly and . High-energy electron beams induce bond breaking and atomic displacement in non-crystalline materials, leading to structural alterations that obscure true atomic arrangements; for example, organic amorphous solids suffer rapid , necessitating low-dose strategies and cryogenic cooling to mitigate damage rates by factors of 10–100. is equally demanding, as amorphous solids often require ultrathin sections (below nm) via milling or , which can introduce artifacts like or bending in beam-sensitive specimens such as . These hurdles demand specialized protocols, including liquid-nitrogen cooling and correlative , to ensure reliable nanoscale visualization. Computational validation of reconstructed images can occasionally aid in assessing these artifacts.

Computational and Modeling Approaches

Computational modeling of amorphous solids relies on simulation techniques that capture the disordered atomic arrangements and without long-range order. (MD) simulations solve Newtonian to evolve atomic positions over time, enabling the study of structural formation through rapid from a high-temperature melt. In these simulations, systems are cooled at rates on the order of 10^9 to 10^14 K/s to mimic glass formation, often using empirical potentials like the Lennard-Jones (LJ) potential to describe pairwise interactions between atoms. A seminal example is the Kob-Andersen binary LJ mixture, which models a supercooled prone to , allowing investigation of relaxation and medium-range order in metallic glasses. Monte Carlo (MC) methods complement MD by facilitating equilibrium sampling in the glassy regime below the temperature (Tg), where kinetic trapping hinders standard simulations. Techniques like swap exchange particle sizes or configurations between replicas at different temperatures, enhancing ergodic exploration and accessing ultra-stable states deep in the supercooled phase. This approach has been applied to two-dimensional glass-forming systems, revealing thermodynamic properties such as vanishing configurational at low temperatures. Density functional theory (DFT) provides accurate electronic structure calculations for small clusters or limited-size models of amorphous solids, treating quantum mechanical effects explicitly. In annealing simulations, DFT optimizes atomic configurations by minimizing energy on surfaces, yielding realistic structures for materials like where classical potentials fall short. For instance, DFT-based relaxation of supercells produces radial distribution functions consistent with experimental observations, highlighting local bonding motifs in disordered networks. Advancements in potentials, particularly neural network-based developed since the 2010s, enable large-scale simulations of amorphous alloys by approximating quantum-accurate energies and forces from training data. These potentials, trained on DFT datasets, surpass traditional empirical models in transferability across phases, allowing MD-like simulations of systems with thousands of atoms, such as amorphous carbon or metallic glasses, to probe structural heterogeneity and dynamics efficiently. Validation of these models against experiments, such as through comparison of simulated radial distribution functions (RDFs) with data, ensures structural fidelity. For example, simulations incorporating bias potentials to refine RDFs have matched large-angle scattering results for amorphous polymers, confirming interatomic correlations and assembly motifs. Similarly, ab initio models of hydrogenated demonstrate that while RDFs provide baseline agreement, complementary vibrational spectra offer stricter benchmarks for local order. Simulations are benchmarked against techniques like or to verify short- and medium-range structures without delving into specific property computations.

Applications and Phenomena

Thin Films and Nanotechnology

Amorphous solids play a pivotal role in technologies and , where their lack of long-range order enables unique properties such as and tunable characteristics at the nanoscale. In s, typically ranging from a few nanometers to hundreds of nanometers in thickness, amorphous materials are deposited to form uniform layers that serve as active components in devices, leveraging their ability to maintain structural homogeneity without grain boundaries. applications extend this to nanostructured composites and films, where amorphous matrices host nanocrystals or other nanostructures, enhancing mechanical and functional performance. One prominent application is in amorphous semiconductors, particularly hydrogenated amorphous silicon (a-Si:H), which is widely used as the absorber layer in thin-film solar cells due to its low-cost deposition and suitable bandgap for photovoltaic conversion. a-Si:H solar cells achieve power conversion efficiencies of approximately 10-15% in laboratory settings, with optimized structures reaching up to 12.71% for ultrathin absorbers around 1 μm thick. This efficiency stems from the material's high absorption coefficient in the , allowing effective light harvesting in films as thin as 300 nm, though challenges like the Staebler-Wronski effect limit long-term stability. Metallic glass thin films, formed by rapid techniques, offer exceptional mechanical properties for microelectromechanical systems (), including high strength-to-weight ratios and elastic limits exceeding 2% strain. These films, often composed of alloys like Zr-Cu-Ni, exhibit yield strengths over 1 GPa while maintaining densities comparable to crystalline metals, making them ideal for robust, lightweight components such as sensors and actuators. Their amorphous structure suppresses motion, providing superior wear resistance and fatigue life compared to polycrystalline counterparts. In , amorphous nano-composites enhance by embedding nanocrystals within an amorphous matrix, creating interfaces that deflect cracks and promote shear banding for improved . For instance, crystalline-amorphous nano-laminates or core-shell structures in metal-glass systems can achieve significantly enhanced compared to pure amorphous metals, by distributing stress and preventing . This design exploits the nanocrystals' ability to pin shear bands in , balancing high strength (often >2 GPa) with enhanced . Key deposition techniques for achieving uniform amorphous thin films include pulsed laser deposition (PLD) and (ALD), which enable precise control over film thickness down to sub-10 nm layers. PLD uses high-energy pulses to ablate targets, producing stoichiometric amorphous films with high density and minimal substrate heating, suitable for complex oxides. ALD, in contrast, employs sequential self-limiting surface reactions for conformal, pinhole-free coatings on high-aspect-ratio nanostructures, ensuring uniformity even on 3D features. These methods are essential for , as they preserve the amorphous state during growth. A major challenge in amorphous thin films is stress-induced crystallization, where intrinsic or extrinsic stresses during deposition or operation promote of crystalline phases, degrading the desired isotropic properties. Compressive stresses exceeding 1 GPa can lower the barrier for in materials like a-Si:H or metallic glasses, leading to and reduced performance in devices; mitigation strategies include alloying or controlled annealing to manage without triggering unwanted ordering.

Superconductivity and Electrical Uses

In amorphous solids, electrical conduction often deviates from crystalline counterparts due to structural disorder, which localizes electron wavefunctions and impedes diffusive transport. This phenomenon, known as , arises when disorder strength exceeds a critical threshold, transforming extended states into localized ones and leading to insulating behavior even at zero temperature. In amorphous semiconductors, where charge carriers occupy localized states near the band edges, conduction proceeds via thermally activated hopping between these sites. At low temperatures, nearest-neighbor hopping dominates, but as temperature decreases further, carriers favor longer-range hops to minimize energy barriers, resulting in (VRH) conduction. The seminal Mott model for VRH in three dimensions describes the DC conductivity \sigma as \sigma = \sigma_0 \exp\left[ -\left( \frac{T_0}{T} \right)^{1/4} \right], where \sigma_0 is a prefactor related to the density of states and wavefunction overlap, T_0 characterizes the localization length and density of states at the Fermi level, and T is temperature; this T^{-1/4} dependence has been experimentally verified in materials like amorphous silicon and chalcogenide glasses. The metal-insulator transition (MIT) in amorphous solids marks the boundary between metallic and insulating regimes, often modeled by percolation theory, which treats conduction as the formation of a continuous network of interconnected metallic regions amid insulating ones. As the metal concentration or disorder varies, the system approaches a percolation threshold where the conductivity exhibits critical scaling behavior, with the correlation length diverging as \xi \sim |p - p_c|^{-\nu}, p being the fraction of metallic bonds and p_c the critical percolation probability. In amorphous alloys like Si-Au or Ge-Au, this transition aligns with the Mott criterion for localization, where the dimensionless parameter k_F l \approx 1 (with k_F the Fermi wavevector and l the mean free path) separates extended from localized states, enabling tunable resistivity over orders of magnitude. Amorphous superconductors, such as molybdenum-germanium alloys (a-MoGe), maintain superconductivity despite disorder, with critical temperatures T_c reaching up to 7 K in thin films, where atomic-scale randomness enhances electron-phonon coupling and pairing compared to crystalline analogs. This disorder-induced enhancement allows persistent superconductivity even near the superconductor-insulator transition, as probed by varying film thickness or composition. Such properties make amorphous solids ideal for superconducting electronics, particularly in Josephson junctions, where uniform thin films serve as electrodes or barriers, enabling high-critical-current densities and reduced quasiparticle tunneling for applications in quantum computing and sensitive detectors. Thin-film devices based on these materials, like SQUID magnetometers, benefit from the lack of grain boundaries, ensuring reproducible Josephson coupling and low noise.

Thermal and Protective Applications

Amorphous silica aerogels, characterized by their highly porous structure composed of interconnected amorphous silica nanoparticles, serve as exceptional thermal insulators in various applications, including building envelopes, cryogenic systems, and aerospace components. Their ultralow thermal conductivity, typically around 0.01 W/m·K at ambient conditions, arises from the minimized solid conduction paths and suppressed gas conduction within the nanoscale pores under or . This property makes them superior to traditional insulators like , enabling efficient heat management in extreme environments while maintaining structural integrity due to the rigid, amorphous network. Amorphous carbon materials, including (DLC) films and soot-derived coatings, play a critical role in barrier applications for high-temperature systems such as engines. DLC coatings, with their amorphous sp³-rich structure, provide robust protection by reflecting heat and reducing substrate temperatures, while exhibiting low conductivity and high oxidation resistance up to 500°C. In turbine blades, these coatings mitigate stresses, extending component lifespan under operating conditions exceeding 1000°C, and their tunable amorphous allows for optimized and . The reusable thermal protection system of the exemplifies the use of amorphous solids in extreme heat shielding, employing low-density tiles fabricated from 99.8% pure amorphous silica fibers. These tiles, with a of approximately 90%, form a rigidized that withstands peak re-entry temperatures up to 1600°C on the orbiter's underside, primarily through radiative heat dissipation and minimal conductive transfer. The amorphous fiber structure ensures low thermal conductivity (around 0.1 W/m·K) and rapid response to thermal gradients, preventing structural failure during . Chalcogenide-based phase-change materials (PCMs), such as Ge-Sb-S-Se-Te alloys, utilize transitions between amorphous and crystalline phases for applications in devices. These materials switch states at controlled temperatures (e.g., near 215°C), enabling high-density through structural reconfiguration with enhanced cycling stability. This amorphous-to-crystalline switching provides reliable performance in , with thermal stability indicated by temperatures around 170–200°C. Amorphous polymers, such as polystyrene, are integral to lightweight radiation shielding composites, where their non-crystalline matrix accommodates high-Z fillers like lead oxide nanoparticles to attenuate gamma rays effectively. In polystyrene-PbO nanocomposites, the amorphous polymer enables uniform dispersion of 52 nm PbO particles at loadings up to 35 wt%, achieving linear attenuation coefficients up to 3.5 times higher than pure polymer at 0.059 MeV, while reducing half-value layer thickness to 0.25 cm. These materials excel in space and medical shielding due to their flexibility, low density (1.03 g/cm³), and enhanced shielding efficiency (e.g., 70% less lead mass than metallic shields) across energies from 0.06 to 1.3 MeV.

Pharmaceutical and Biological Uses

Amorphous solids play a crucial role in pharmaceutical applications due to their enhanced and compared to crystalline counterparts, particularly for poorly water-soluble drugs. For instance, amorphous indomethacin exhibits significantly higher aqueous —up to several orders of magnitude greater than its crystalline form—leading to improved rates and oral . However, this thermodynamic instability often results in recrystallization, or , which can revert the material to a less soluble crystalline state, posing challenges for long-term storage and efficacy. To mitigate these stability issues, amorphous solid dispersions (ASDs) are commonly formulated by dispersing the drug within a polymer matrix, which inhibits devitrification through antiplasticization effects and molecular interactions that reduce molecular mobility. Polymers such as (PVP) or hydroxypropyl methylcellulose (HPMC) are frequently used as stabilizers, enhancing physical stability by increasing the temperature and preventing during storage or dissolution. Common preparation methods include spray-drying, where a drug-polymer is atomized and rapidly dried to form amorphous particles, and hot-melt extrusion, which involves melting and extruding the mixture to produce stable ASDs with controlled release profiles. These techniques have enabled the approval of numerous ASD-based drug products by the U.S. , with 48 such formulations authorized between 2012 and 2023 as of 2023, demonstrating their growing clinical impact. In biological contexts, amorphous solids appear naturally as precursors in processes, notably amorphous calcium phosphate (ACP), which serves as an intermediate phase in formation. ACP nanoparticles are transiently deposited in the of developing , providing a flexible, hydrated precursor that transforms into crystalline under physiological conditions, facilitating rapid mineralization and structural adaptation. This amorphous phase's high solubility and reactivity enable efficient ion transport and integration into scaffolds, underscoring its role in skeletal development and repair.

Occurrence in Nature

Amorphous solids occur widely in natural geological processes, particularly where rapid cooling prevents . Obsidian, a type of , forms when silica-rich lava from viscous eruptions cools extremely quickly, resulting in an amorphous structure with no crystalline minerals. This natural glass is composed primarily of silica (SiO₂) and minor oxides, exhibiting a brittle texture and ideal for sharp edges. Humans have exploited 's properties since prehistoric times, crafting tools, weapons, and arrowheads from its razor-sharp flakes due to its homogeneous composition and ease of flaking. Impact glasses, such as tektites, represent another geological manifestation of amorphous solids, generated during collisions with Earth's surface. These events produce intense heat and pressure, melting target rocks and ejecting molten material that solidifies into glassy droplets or fragments upon atmospheric re-entry, forming non-crystalline structures rich in silica and alumina. Tektites are found in strewn fields across continents, like the Australasian field covering millions of square kilometers, and their aerodynamic shapes confirm high-speed formation. In biological systems, amorphous silica manifests as opal-A, a hydrated form (SiO₂·nH₂O) biosynthesized by organisms for structural support. Diatoms, unicellular , construct intricate frustules—cell walls—from this non-crystalline silica, enabling their role in productivity and the . Similarly, certain sponges, such as demosponges, incorporate opal-A into spicules and skeletal frameworks, contributing significantly to silica and fluxes. These biogenic structures dissolve more readily than crystalline silica, influencing availability in environments. Soils host amorphous components essential for fertility and structure, including and allophane. , derived from decomposed , are complex, amorphous macromolecules that bind soil particles, enhance water retention, and chelate nutrients like iron and . These dark, polyelectrolyte-like materials form through microbial processes and persist in amorphous states, resisting . Allophane, a short-range ordered (Al₂O₃·SiO₂·nH₂O), occurs in weathered volcanic soils and andisols, acting as a gel-like adsorbent for anions and cations due to its amorphous nature and high surface area. Atmospheric aerosols include amorphous sulfate phases that influence and cloud processes. Sulfate aerosols, often as or particles, can adopt semi-solid or glassy states at low temperatures and humidities, particularly when mixed with organics, increasing and altering particle dynamics. These glassy sulfates in the and promote heterogeneous , affecting formation and potentially amplifying climate warming by altering precipitation efficiency. Volcanic injections of further enhance sulfate glass formation, leading to temporary cooling via enhanced aerosol scattering.

Phase Behavior

Structural Relaxation

Structural relaxation in amorphous solids refers to the time-dependent evolution of their atomic or molecular structure toward a more stable configuration below the glass transition temperature (Tg). This process occurs as the material, initially in a non-equilibrium state formed by rapid cooling, gradually minimizes its through cooperative rearrangements. Unlike crystalline solids, amorphous materials lack long-range order, leading to heterogeneous dynamics where local regions relax at different rates, resulting in overall structural changes over extended periods. Physical aging, a key manifestation of structural relaxation, involves densification and a decrease in as the amorphous solid approaches its supercooled state. During aging, the volume contracts due to the reconfiguration of bonds, such as the reduction of less homopolar linkages in favor of more energetically favorable heteropolar ones, leading to increased packing . Concurrently, the decreases as the system sheds excess stored during , with losses on the order of 0.5–1 per atom in some metallic . These changes unfold over broad time scales, from seconds in accelerated conditions near to years at ambient temperatures, reflecting the sluggish inherent to glassy . The relaxation kinetics during aging are often described by the , also known as the Kohlrausch-Williams-Watts (KWW) form: \phi(t) = \exp\left[-\left(\frac{t}{\tau}\right)^\beta\right] where \phi(t) is the relaxation function, \tau is the characteristic relaxation time, and \beta (0 < \beta < 1) is the stretching exponent that captures the non-exponential, distributed nature of the process in disordered systems. This form, originally proposed by Kohlrausch in for charge decay in insulators, empirically fits structural relaxation in , with typical \beta values around 0.3–0.6 near Tg, indicating broad heterogeneity in relaxation times. The stretched exponential arises from trapping models where diffusive motion in disordered landscapes leads to subdiffusive behavior, providing a for the cooperative yet spatially varying rearrangements in amorphous solids. Under applied , physical aging exhibits non-linear effects, where the relaxation accelerates or decelerates depending on the magnitude and direction relative to the material's aging history. In nonlinear viscoelastic models, perturbs the energy landscape, enhancing local mobility and causing faster structural compared to isothermal aging without load, as seen in temperature-jump experiments on glasses. This nonlinearity is quantified through material time scaling, where the effective relaxation time shortens under tensile but may lengthen under , leading to asymmetric responses that deviate from linear superposition principles. Such effects are critical in modeling the viscoelastic behavior of amorphous polymers under mechanical loading. The β-relaxation serves as a secondary and precursor to the primary α-relaxation in amorphous , initiating localized motions that facilitate the larger-scale structural changes of the α . Observed below , the JG relaxation involves short-range collective atomic rearrangements, such as bonding switches in metallic glasses that increase local density by favoring solute-solvent interactions over solvent-solvent pairs, with activation energies around 25–30 RTg. This β , universal across glassy materials, acts as a dynamic heterogeneity driver, to the α relaxation by providing the initial fluctuations necessary for cooperative in the viscous regime. These relaxation processes have significant implications for the mechanical of amorphous solids, as aging-induced densification and reduction enhance rigidity but can also introduce . For instance, increased structural relaxation in polymer glasses correlates with higher and , improving resistance to deformation but reducing tensile strength in compacted forms due to diminished irreversible work absorption. In metallic glasses, JG-mediated relaxation modulates the transition from anelastic to deformation, influencing stability and overall durability under load. Thus, controlling structural relaxation is essential for tailoring mechanical performance in applications like coatings and structural materials.

Polyamorphism and Transitions

Polyamorphism refers to the existence of multiple distinct amorphous phases in a single-component material, differing in density and structure, much like polymorphism in crystals. Recent studies have identified intermediate-density amorphous ices (MDA), suggesting a possible continuum of amorphous structures between low-density and high-density forms under certain pressure and temperature conditions. In the case of water, two prominent amorphous forms are low-density amorphous ice (LDA), with a density of approximately 0.94 g/cm³ at ambient pressure, and high-density amorphous ice (HDA), with a density of about 1.17 g/cm³. LDA features a tetrahedral hydrogen-bonded network similar to ice Ih, while HDA exhibits a more collapsed structure with increased coordination numbers around 5. These phases form under varying pressure conditions: LDA through vapor deposition or hyperquenching of liquid water, and HDA via compression of crystalline ice Ih at low temperatures (e.g., 77 K) to around 1 GPa. The transition between LDA and HDA is a first-order-like process, occurring discontinuously at pressures near 0.6 GPa and temperatures around 77 , accompanied by a ~20% volume change and upon reversal. This polyamorphic switch highlights how external pressure can induce structural reorganization in amorphous solids without . A related phenomenon is the very high-density amorphous ice (VHDA), formed by annealing HDA at pressures above 1 GPa and temperatures near 125 , featuring even higher coordination (around 6) and density. These pressure-induced transformations underscore the sensitivity of amorphous phases to thermodynamic conditions, enabling distinct metastable states. The liquid-liquid transition (LLT) hypothesis posits that in the supercooled regime of (below 235 K at ), a occurs between low-density liquid (LDL) and high-density liquid (HDL) phases, mirroring the LDA-HDA polyamorphism upon . Proposed in simulations showing a critical point at ~220 K and 0.1 GPa, with recent 2025 estimates refining it to approximately 200 K and 0.13 GPa, the LLT explains anomalies like maxima and divergences in supercooled . Experimental emerged from ultrafast heating of HDA at 205 K under 2.5–3.5 kbar, where revealed HDL formation followed by LDL domain growth during decompression, confirming a discontinuous structural shift distinct from . This supports the idea that amorphous polyamorphism reflects underlying liquid phase behavior in the deeply supercooled "no-man's land." In amorphous solids derived from undercooled melts, crystallization proceeds via nucleation and growth pathways, where nucleation—the formation of critical embryos—often decouples from subsequent crystal growth due to differing temperature dependencies. For glass-formers like metallic alloys (e.g., Zr-Ti-Cu-Ni-Be) or pharmaceuticals (e.g., ibuprofen), nucleation rates peak near or slightly above the glass transition temperature (T_g), driven by thermodynamic undercooling (ΔT = T_m - T, where T_m is the melting point), while growth maxima occur closer to T_m where mobility is higher. Heterogeneous nucleation dominates at surfaces or impurities, leading to polymorphic sequences where a metastable crystal form nucleates first, followed by transformation to the stable phase; for instance, in l-arabitol, nucleation maximizes at 5°C (T_g ≈ -14°C), while growth peaks between 60–95°C (T_m ≈ 101°C). These pathways determine the stability of amorphous materials, with deep undercooling favoring glass formation over rapid crystallization. Metastable phase diagrams for glass-formers map the undercooled , amorphous, and crystalline regions, incorporating polyamorphic boundaries and avoiding true lines. In , such diagrams depict the supercooled extending to ~235 , with the LLT critical point connecting LDL/HDL to LDA/HDA upon cooling, and a "no-man's land" below 227 where rapid hinders access. For general glass-formers, these diagrams highlight the glass-forming range bounded by T_g (isothermal ) and T_m, with pressure axes revealing polyamorphic transitions; for example, in metallic systems, metastable extensions predict amorphous matrix compositions up to 28% solute before intervenes. These diagrams guide processing conditions to stabilize amorphous phases. High-pressure neutron scattering provides direct structural evidence for polyamorphism, revealing distinct pair correlation functions for LDA and HDA. Early studies at ~1 GPa showed HDA's first diffraction peak shifting to higher (indicating shorter O-O distances ~2.8 vs. 2.95 in LDA), confirming the collapsed network without crystalline order. More recent inelastic on HDA under (e.g., 0.2–2 GPa) tracked the LDA-HDA , observing mode softening and jumps consistent with a first-order change. These techniques validate the polyamorphic phases and their pressure-driven interconversions in amorphous solids.

References

  1. [1]
    Amorphous - CES Information Guide - Materials Science Engineering
    Amorphous Structure An amorphous structure has no organization (not a crystalline structure), and the atomic structure resembles that of a liquid.
  2. [2]
    Crystalline and Amorphous Solids
    Amorphous solids have two characteristic properties. When cleaved or broken ... Crystals have sharp, well-defined melting points; amorphous solids do not.
  3. [3]
    Amorphous materials - UCSB Physics
    Amorphous solids are comprised of particles (atoms, grains, bubbles, molecules) arranged so that the locations of their centers of mass are disordered.
  4. [4]
    Amorphous Metals - MRSEC Education Group
    The atoms in an amorphous material are not arranged in any ordered structure, rather they have a tightly-packed, but random arrangement. Amorphous materials ...Missing: definition | Show results with:definition
  5. [5]
    AMORPHOUS PHARMACEUTICAL SOLIDS - PMC - NIH
    Amorphous forms are, by definition, non-crystalline materials which possess no long-range order. Their structure can be thought of as being similar to that of a ...
  6. [6]
    On Structure and Properties of Amorphous Materials - PMC
    Mechanical, optical, magnetic and electronic properties of amorphous materials hold great promise towards current and emergent technologies.
  7. [7]
    Amorphous Semiconductors - Materials Science Wiki (DG)
    Jan 9, 2007 · An amorphous solid is defined as one that does not maintain long-range translational symmetry or has only short-range order, it does not have ...Bond States · Structure · Band Diagram
  8. [8]
    The Solid State of Matter – Chemistry - UH Pressbooks
    The resulting materials are called amorphous solids or noncrystalline solids (or, sometimes, glasses). The particles of such solids lack an ordered internal ...The Solid State Of Matter · Learning Objectives · Properties Of Solids<|control11|><|separator|>
  9. [9]
    [PDF] 1. Introduction Amorphous solids exhibit many universal ...
    Amorphous solids exhibit many universal phenomena that are absent in crystalline solids. One such example is the low energy excitations [1], where, at low ...
  10. [10]
    Review: understanding the properties of amorphous materials with ...
    May 22, 2023 · We will use the definition that an amorphous material is one which has no long-range structural order. This lack of order has important ...Introduction to amorphous... · Molecular dynamics... · Applications · Conclusion
  11. [11]
    Is the Future of Materials Amorphous? Challenges and ...
    Amorphous materials are typically defined as lacking long-range order, which means their atomic structure does not produce the well-defined diffraction patterns ...
  12. [12]
    12.1: Crystalline and Amorphous Solids - Chemistry LibreTexts
    Apr 12, 2023 · Crystalline solids have well-defined edges and faces, diffract x-rays, and tend to have sharp melting points. In contrast, amorphous solids have ...
  13. [13]
    Fact or Fiction?: Glass Is a (Supercooled) Liquid | Scientific American
    Feb 22, 2007 · "Amorphous means it doesn't have that long-range order," Ediger says. With a "solid—if you grab it, it holds its shape," he adds. When glass ...
  14. [14]
    Testing the paradigm of an ideal glass transition: Dynamics of an ...
    Dec 21, 2018 · INTRODUCTION. Conventional glass is a nonequilibrium material made by rapid cooling from the equilibrium or the supercooled liquid states.
  15. [15]
    Amorphous - Etymology, Origin & Meaning
    Originating from Greek amorphos ("without form") via Modern Latin amorphus (1731), amorphous means shapeless or lacking a determined form.Missing: solid | Show results with:solid
  16. [16]
    The Project Gutenberg eBook of Micrographia, by Robert Hooke
    May 23, 2021 · Some physiological descriptions of minute bodies made by magnifying glasses with observations and inquiries thereupon.
  17. [17]
    Amorphous solid | Properties, Structure & Examples - Britannica
    Amorphous solid, any noncrystalline solid in which the atoms and molecules are not organized in a definite lattice pattern. Such solids include glass, plastic, ...Preparation of amorphous solids · Properties of oxide glasses · Transparent Glasses
  18. [18]
    Page:EB1911 - Volume 06.djvu/80 - Wikisource, the free online library
    The results of Berzelius were greatly extended by Hermann Kopp, who ... amorphous forms of carbon having the value about 5.6 at 1000°, and silicon ...
  19. [19]
    Two Centuries of Glass Research: Historical Trends, Current Status ...
    Jul 28, 2014 · In this article, we analyze the number of research papers and patents related to glassy and amorphous materials in the published literature.
  20. [20]
    A Brief Scientific History of Glass - Smithsonian Magazine
    Nov 24, 2021 · Based on some beautiful, well-preserved glass artifacts dating from around 1500 BCE, Egypt was favored at first.
  21. [21]
    The ancient origins of glass | Knowable Magazine
    Nov 18, 2021 · Based on some beautiful, well-preserved glass artifacts dating from around 1500 BCE, Egypt was favored at first. But by the 1980s, researchers ...
  22. [22]
    Crafting the microworld: how Robert Hooke constructed knowledge ...
    Dec 16, 2015 · His Micrographia is justifiably famous for its detailed engravings, which communicated Hooke's observations ... Gums, Resins, Salts ...
  23. [23]
    Jöns Jakob Berzelius | Science History Institute
    An avid and methodical experimenter, Jöns Jakob Berzelius (1779–1848) conducted pioneering experiments in electrochemistry and established the law of constant ...
  24. [24]
    Thomas Graham | Chemistry Pioneer, Physicist, Experimenter
    Sep 26, 2025 · The scientific study of colloids dates from the early 19th century. Among the first notable investigations was that of the British botanist ...Missing: amorphous | Show results with:amorphous
  25. [25]
    Colloid - an overview | ScienceDirect Topics
    In 1861, Thomas Graham first introduced the word “colloid” to define a peculiar aggregation form of substances [1]. In the past ∼157 years, colloidal ...
  26. [26]
    Existence of multiple critical cooling rates which generate different ...
    Mar 22, 2019 · The kinetic concept of glass formation was developed by Gustav Tammann 85 years ago. ... cooling rate at which the liquid can form a glass. Fig. 1 ...
  27. [27]
    [PDF] The glassy state of matter: Its definition and ultimate fate - LaMaV
    Jun 8, 2017 · Glasses (G) exist below the glass transition temperature, Tg. They are thermodynamically unstable and spontaneously relax toward the supercooled ...Missing: undercooling | Show results with:undercooling
  28. [28]
    THE ATOMIC ARRANGEMENT IN GLASS - ACS Publications
    W. H. Zachariasen. ACS Legacy Archive. Open PDF. Journal of the American Chemical Society. Cite this: J. Am. Chem. Soc. 1932, 54, 10, 3841–3851. Click to copy ...Missing: URL | Show results with:URL
  29. [29]
    Non-crystalline Structure in Solidified Gold–Silicon Alloys - Nature
    Nature volume 187, pages 869–870 (1960)Cite this article. 16k ... Low-temperature glow discharge plasma oxidation of the Zr48Cu36Al8Ag8 bulk metallic glass.Missing: paper | Show results with:paper
  30. [30]
    40 Years Trajectory of Amorphous Semiconductor Research
    Aug 6, 2025 · A review is given on a research trajectory of amorphous and microcrystalline semiconductors and their device applications proceeded since 1970.
  31. [31]
    Computer-simulation study of the thermal conductivity of amorphous ...
    Aug 1, 1996 · We study by computer simulation an amorphous solid modeled as a percolating cluster of atoms on the fcc lattice. The atoms interact by ...Missing: computational | Show results with:computational
  32. [32]
    Customized design of amorphous solids by generative deep learning
    A deep learning architecture combining generation and prediction for design of amorphous solids is developed. It enables automated generation of compositions ...Missing: 2020s | Show results with:2020s
  33. [33]
    Application of Machine Learning in Amorphous Alloys - PMC
    Apr 13, 2025 · This review focuses on four key applications of ML in amorphous alloys: (1) prediction of amorphous alloy phases, (2) prediction of amorphous composite phases,Missing: 2020s | Show results with:2020s
  34. [34]
    Radial distribution function and structural relaxation in amorphous ...
    Dec 15, 1981 · A method of interpreting radial distribution functions (RDF) of amorphous metals is proposed in which the role of the local atomic structure is emphasized.
  35. [35]
    Pressure-induced amorphization and polyamorphism
    Pressure-induced amorphization (PIA) is a phenomenon that involves an abrupt transition between a crystalline material and an amorphous solid through ...
  36. [36]
    [PDF] Melt-spinning technique for preparation of metallic glasses
    To achieve cooling rates of this magnitude, all rapid quenching techniques are based on the principle of bringing the melt in contact with a highly conducting ...
  37. [37]
    [PDF] Greer - Rutgers University, GLASS |
    They have excited interest in metallic glasses based on light metals and have proved to be the basis for further sig- nificant developments in ultrafine-scale.
  38. [38]
    Effect of deposition rate on micromorphology analyses and optical ...
    May 10, 2025 · In nanomaterials in the form of amorphous composite thin films by sputtering, two key elements are always effective in the role of the final ...
  39. [39]
    Ultrathin amorphous carbon films synthesized by filtered cathodic ...
    Jun 25, 2018 · Contemporary magnetic heads and hard disks are coated with a thin a-C film deposited by plasma-enhanced chemical vapor deposition or filtered ...
  40. [40]
    Amorphisation of solids by ion implantation - ScienceDirect.com
    Ion implantation of many solids frequently results in phase changes to a greater or lesser extent. The nature and extent of the phase change depends upon ...Missing: state | Show results with:state
  41. [41]
    Amorphization of Zr–Al solid solutions under mechanical alloying at ...
    May 1, 1999 · The effects of temperature on the amorphization of Zr–Al solid solutions have been investigated by ball milling Zr100−xAlx powder blends at
  42. [42]
    Mechanical alloying and milling - ScienceDirect.com
    Mechanical alloying (MA) is a solid-state powder processng technique involving repeated welding, fracturing, and rewelding of powder particles in a high-energy ...
  43. [43]
    Nanomaterial by Sol‐Gel Method: Synthesis and Application
    Dec 24, 2021 · The sol-gel method can be used in the process of making ceramics as a molding material and can be used as an intermediate between thin films ...Advantages of Sol-Gel Method · Sol-Gel Process Steps · Application of the...
  44. [44]
    Pressure-induced amorphization and existence of molecular ... - PNAS
    Apr 3, 2020 · By using experimental high-pressure techniques, we observe a reversible transition between different amorphous states around 26 GPa.
  45. [45]
    Dependence of glass forming ability on liquid fragility - AIP Publishing
    Nov 1, 2010 · It is found that increasing fragility thermodynamically favors glass formation, remarkably contrasting with the fragility effect from the kinetic point of view.
  46. [46]
    Correlation between fragility and glass-forming ability of metallic alloys
    Sep 6, 2007 · This GFA parameter, which increases with a decrease in the critical cooling rate, is a function of the reduced glass transition temperature T r ...
  47. [47]
    First-principles study on the specific heat jump in the glass transition ...
    Sep 15, 2023 · In this study, first-principles molecular-dynamics simulations are used to describe the glass transition of silica glass.
  48. [48]
    Structural Relaxation Rate and Aging in Amorphous Solids
    Mar 17, 2023 · Parameters for enthalpy relaxation in and below the glass transition temp. ... enthalpy line can indeed be reached for aging 15 K below Tg. The ...
  49. [49]
    Enthalpy Relaxation - an overview | ScienceDirect Topics
    Enthalpy relaxation refers to the release of degrees of freedom upon heating a material that has been aged below its glass transition temperature, with its ...
  50. [50]
    A structural signature of liquid fragility | Nature Communications
    Aug 6, 2014 · We define a structural fragility index, γ, as the mismatch between the value for S(q1) in the liquid when extrapolated to Tg and the value ...
  51. [51]
    Poisson's Ratio and Intrinsic Plasticity of Metallic Glasses |
    Jun 30, 2008 · Assuming a sphere-shaped STZ, ξ c = 1.782 corresponds to a Poisson's ratio ν c ≈ 0.415 ⁠. For comparison, the critical Poisson's ratio ...
  52. [52]
    Understanding amorphous phase-change materials from the ...
    May 26, 2010 · At the heart of the rigidity concept is the identification of relevant interatomic forces between atoms in a manner similar to what Maxwell ...
  53. [53]
    [PDF] Mechanical behavior of amorphous alloys
    Isotropic materials have two independent elastic constants, commonly taken to be Young's modulus (E) and Poisson's ratio (m). From a more fundamental point of ...
  54. [54]
    Experimental characterization of shear transformation zones ... - PNAS
    We report experimental characterization of shear transformation zones (STZs) for plastic flow of bulk metallic glasses (BMGs) based on a newly developed ...
  55. [55]
    Ductile to brittle transition of fracture of a Zr-based bulk metallic glass
    The former two factors (temperature and free volume) have been confirmed by uniaxial loading experiments, while the third one (strain rate) has been rarely shed ...
  56. [56]
    Mechanical glass transition revealed by the fracture toughness of ...
    Aug 16, 2018 · We show that the notch fracture toughness of metallic glasses exhibits an abrupt toughening transition as a function of a well-controlled fictive temperature.
  57. [57]
    Review Article Creep in amorphous metals - ScienceDirect.com
    This paper reviews the work on creep behavior of amorphous metals. There have been, over the past several years, a few reviews of the mechanical behavior of ...
  58. [58]
    Fatigue and fracture behavior of bulk metallic glasses and their ...
    Starting from 1975, the fatigue behavior of metallic glasses was reported [67], [68]. In general, the samples of these metallic glasses are ribbons or wires.
  59. [59]
    Low-temperature thermal conductivity and acoustic attenuation in ...
    Oct 30, 2002 · All measurements published on the low-temperature thermal conductivity and acoustic attenuation in these solids have been reviewed, on a total of over 60 ...Missing: TLS | Show results with:TLS<|separator|>
  60. [60]
    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 ...
  61. [61]
    High frequency sound and the boson peak in amorphous silica
    We show that the low frequency part of the boson peak is mainly due to the elastic scattering of transverse acoustic modes with frequencies around 1 THz. The ...
  62. [62]
    Internal-friction peak in amorphous selenium at low temperature
    The discovery of a large internal-friction peak in amorphous selenium around 30K at 160 MHz is reported. The peak appears on the attenuation of ultrasonic shear ...
  63. [63]
    The rise of the X-ray atomic pair distribution function method - Journals
    Apr 29, 2019 · The atomic pair distribution function (PDF) technique is a powerful approach to gain quantitative insight into the structure of materials.Missing: halos | Show results with:halos
  64. [64]
    Structural Analysis of Molecular Materials Using the Pair Distribution ...
    Nov 17, 2021 · This is a review of atomic pair distribution function (PDF) analysis as applied to the study of molecular materials.
  65. [65]
    Analysis of Amorphous and Nanocrystalline Solids from Their X-Ray ...
    The observation of X-ray amorphous powder patterns may indicate the presence of amorphous, glassy or disordered nanocrystalline material in the sample. Rietveld ...
  66. [66]
    [PDF] Neutron PDF for Local Structure Studies
    Non crystalline materials: Amorphous solids and polymers. ▫ Disordered materials: The interesting properties are often governed by the defects or local.
  67. [67]
    [PDF] THE CHEMICALLY-SPECIFIC STRUCTURE OF AN AMORPHOUS ...
    mechanism is the change in the neutron scattering length as a function of the isotope. The most fundamental difference between neutron and x-ray diffraction is ...
  68. [68]
    Illustrated formalisms for total scattering data: a guide for new ... - NIH
    The total scattering method is the simultaneous study of both the real- and reciprocal-space representations of diffraction data. While conventional Bragg- ...
  69. [69]
    Neutron total scattering method: Simultaneous determination of long ...
    Neutron total scattering provides simultaneous information about long-range order through the Bragg peaks and short-range order through the diffuse ...
  70. [70]
    Compositional dependence of local atomic structures in amorphous ...
    Dec 29, 2006 · Selected area electron diffraction (SAED) patterns were taken using a transmission electron microscope (TEM) equipped with an energy filter.
  71. [71]
    Local structure analysis of amorphous materials by angstrom-beam ...
    Dec 15, 2020 · The angstrom-beam electron diffraction method was used to obtain the local structure information of amorphous materials at a sub-nanometre scale.
  72. [72]
    Review Pair Distribution Function Obtained from Electron Diffraction
    Feb 3, 2021 · Pair distribution function from electron diffraction (ePDF) is a remarkable technique capable of elucidating the atomic arrangement of ...Missing: halos | Show results with:halos
  73. [73]
    Engineering medium-range order and polyamorphism in a ... - Nature
    Sep 27, 2019 · Simultaneous small-angle/wide-angle X-ray scattering (SAXS/WAXS) reveal a 'solidification-like' structure transformation of the ED Ni82P18 alloy ...
  74. [74]
    Wide angle X-ray scattering and atomic force microscopy studies of ...
    Summary. The WAXS technique has been used to characterize short and medium range ordering in the amorphous In40Se60 and In34Se66 films. The performed ...
  75. [75]
    Bringing diffuse X-ray scattering into focus - PMC - NIH
    The inherent limitations of Bragg analysis, however, which only reveals the mean structure, have given way to a surge in interest in diffuse scattering, which ...
  76. [76]
    On Structure and Properties of Amorphous Materials - MDPI
    Mechanical, optical, magnetic and electronic properties of amorphous materials hold great promise towards current and emergent technologies.
  77. [77]
    Medium-range order in amorphous silicon measured by fluctuation ...
    We find that every sample of amorphous silicon and germanium we have investigated, regardless of deposition method or hydrogen content, is rich in medium-range ...Missing: solids | Show results with:solids
  78. [78]
    Fluctuation microscopy: a probe of medium range order - IOPscience
    Oct 10, 2005 · Fluctuation microscopy is a hybrid diffraction-imaging technique that detects medium range order in amorphous materials by examining spatial fluctuations in ...
  79. [79]
    Scanning tunneling microscopy of hydrogenated amorphous silicon
    ... surface properties are correlated with bulk properties and local imperfections of as-grown surfaces can indicate the formation of bulk defects, respectively.
  80. [80]
    Scanning tunneling microscopy of defects induced by carbon ...
    Most of the defects observed by STM seem to result from interstitial migrations and lattice rearrangements following disorder generation by the cascade. It is ...
  81. [81]
    Cryogenic 4D-STEM analysis of an amorphous-crystalline polymer ...
    Mar 18, 2022 · We demonstrate a cryogenic 4D-STEM technique using a combination of amorphous radial profile mapping and correlative crystalline growth processing methods.
  82. [82]
    CryoTEM as an Advanced Analytical Tool for Materials Chemists
    Jun 30, 2017 · We utilize cryogenic transmission electron microscopy (cryoTEM), which can track the structural evolution of materials in solution with nanometer spatial ...
  83. [83]
    Reducing the effects of radiation damage in cryo-EM using liquid ...
    Apr 22, 2025 · Cooling the specimen to liquid-nitrogen temperature (≃80 K) is the only successful method to date of reducing the effects of radiation damage ...
  84. [84]
    Challenges in making ideal cryo-EM samples - ScienceDirect
    Achieving better control of the sample thickness · Ice-free cryo-EM sample preparation · Deposition of individual, pL droplets · Writing or streaking · Concept of ...
  85. [85]
    [PDF] Is the Future of Materials Amorphous? Challenges and ... - arXiv
    The preparation of molecular glasses primarily involves three techniques: Liquid-quenching methods, spin-coating, and physical vapor deposition (PVD). Liquid- ...
  86. [86]
    Zero-temperature glass transition in two dimensions - Nature
    Apr 3, 2019 · Here, we develop Monte Carlo methods for two-dimensional glass-forming liquids that allow us to access equilibrium states at sufficiently low ...Missing: Tg | Show results with:Tg
  87. [87]
    [PDF] Numerical investigation of the equilibrium Kauzmann transition in a ...
    Jul 4, 2025 · While conventional molecular dynamics simulations fail well above Tg, the swap Monte. Carlo algorithm can reach below Tg. Here we combine ...
  88. [88]
    Optimal atomic structure of amorphous silicon obtained from density ...
    In order to improve on this approach, we have carried out annealing simulations with a density functional theory (DFT) description of the electronic degrees of ...
  89. [89]
    [PDF] A DFT based study of the low-energy electronic structures and ...
    Häkkinen and Landman6 have investigated the low-energy structures of gold clusters and their anions for n = 2 to 10 atoms using the density-functional theory ( ...
  90. [90]
    A systematic approach to generating accurate neural network ...
    Apr 14, 2021 · Here we propose a self-consistent approach that is based on crystal structure prediction formalism and is guided by unsupervised data analysis.
  91. [91]
    Applications and training sets of machine learning potentials
    This review reports on the selection of training sets for machine learning potentials tailored to their specific applications.<|separator|>
  92. [92]
    Molecular dynamics (MD) simulations and large-angle X-ray ...
    The results validate our earlier finding that MD analysis with implementation of a biasing potential for the RDFs can provide quantitative information on the ...
  93. [93]
    [PDF] Modeling of disordered materials: radial distribution function ... - arXiv
    As molecular dynamics is increasingly used to characterize non-crystalline materials, it is crucial to verify that the numerical model is accurate enough, ...
  94. [94]
    Thin film metallic glasses: Properties, applications and future
    Jan 15, 2020 · In this article, we provide a perspective overview of the iconic properties, recent application-oriented research, and future ...
  95. [95]
    Efficient amorphous silicon solar cells: characterization, optimization ...
    The optimized amorphous silicon solar cell achieved a maximum short-circuit current density of 15.32 mA/cm² and an energy conversion efficiency of 11.3%.Missing: seminal | Show results with:seminal
  96. [96]
    Numerical Design of Ultrathin Hydrogenated Amorphous Silicon ...
    Aug 14, 2021 · The optimized structure of the ultrathin a-Si:H-based solar cell gives a conversion efficiency of 12.71% for a thickness of 1 μm and a bulk ...Missing: seminal | Show results with:seminal<|control11|><|separator|>
  97. [97]
    Thin film metallic glasses as new MEMS materials - ResearchGate
    Aug 5, 2025 · The TFMGs exhibit high strength and high elastic limit at room temperature and low viscosity (under 1013 Pa.s) at supercooled liquid region ( ...Missing: seminal | Show results with:seminal
  98. [98]
    Enhancing strength and ductility via crystalline-amorphous ... - Science
    Mar 9, 2022 · Crystalline-amorphous composite have the potential to achieve high strength and high ductility through manipulation of their microstructures.
  99. [99]
    Insights into the Fracture Mechanisms and Strength of Amorphous ...
    May 12, 2006 · We find that fracture in the nanocomposites occurs intergrain, in the embedding matrix, and so the ideal strength is similar to that of t a - C ...Missing: toughness seminal
  100. [100]
    Pulsed Laser and Atomic Layer Deposition of CMOS-Compatible ...
    Jul 2, 2025 · PLD and ALD offer distinct advantages: PLD enables the formation of high-density films, while ALD allows for conformal deposition on complex 3D ...Pld Method · Figure 3 · Ald MethodMissing: seminal | Show results with:seminal
  101. [101]
    Stress induced crystallization of amorphous materials and ...
    Aug 7, 2025 · It was observed that a tensile stress introduced to an amorphous material can enhance crystallization which eventually serves as an important ...Missing: seminal | Show results with:seminal
  102. [102]
    Stress induced crystallization of hydrogenated amorphous silicon
    Apr 2, 2009 · These results highlight the potential of thermal stress as a means to alter the crystallization in thin film architectures and suggest that even ...
  103. [103]
    Josephson junctions based on amorphous MoGe: prospects for use ...
    Oct 19, 2025 · The results obtained imply that Josephson tunnel junctions based on amorphous superconductors are promising candidates for use in ...
  104. [104]
    Microscopic revelation of the solid–gas coupling and Knudsen effect ...
    Dec 5, 2022 · In this work, molecular dynamics simulations were performed on porous silica aerogel systems to determine their thermal conductivities directly.
  105. [105]
    [PDF] Enabling Martian habitability with silica aerogel via the solid-state ...
    Jul 15, 2019 · However, the thermal conductivity of solid silica (0.8–1.6 W m−1 K−1). (ref. 16) is too high to allow a strong warming effect. Silica aerogels, ...
  106. [106]
    Atomic coordination dictates vibrational characteristics and thermal ...
    Apr 4, 2022 · We discuss the role of atomic coordination in dictating the vibrational characteristics and thermal conductivity in amorphous carbon.
  107. [107]
    Effect of Nitrogen Doping and Temperature on Mechanical Durability ...
    Jul 11, 2018 · Thermal Barrier Coatings for Gas-Turbine Engine Applications. ... Bias-graded deposition of diamond-like carbon for tribological applications.
  108. [108]
    [PDF] The Characterization of the Selected Materials for Space Shuttle
    The FiRST tiles are made of a low-density, high purity silica 99.8-percent amorphous fiber insulation that is made rigid by ceramic bonding resulting in 90 ...
  109. [109]
    [PDF] Space Shuttle Orbiter Thermal rotectlon System Design and Flight ...
    ... fibrous, silica batting made up of high-purity silica, 99.8-percent amorphous silica fibers. This battingis sandwiched betweon an outerwoven silica high-.
  110. [110]
    Phase-change materials based on amorphous equichalcogenides
    Feb 18, 2023 · Chalcogenide phase-change materials (PCMs) are known for their unique behavior during switching between the amorphous and crystalline states.
  111. [111]
    Unravelling the amorphous structure and crystallization mechanism ...
    Feb 3, 2024 · Phase change memory (PCM) is a non-volatile memory technology which exploits the rapid and reversible switching capabilities of phase change ...
  112. [112]
    Enhanced radiation shielding efficiency of polystyrene ... - Nature
    Aug 28, 2024 · This article provides key challenges and opportunities in polymer nanocomposites for radiation shielding and reviews the current state-of-the- ...Material And Methods · Pbo Nanoparticles Synthesis · Linear Attenuation...
  113. [113]
    Development of PMMA based polymer composite incorporating WO ...
    Jul 27, 2025 · In this study, we investigate a novel composite material combining PMMA with tungsten oxide (WO3), a high-atomic-number filler known to enhance ...
  114. [114]
    The Case of Indomethacin - PMC - NIH
    However, the solubility of co-ground amorphous indomethacin in water and phosphate buffer was very high and the amount of co-ground powder needed to conduct ...
  115. [115]
    Preparation of amorphous indomethacin nanoparticles by aqueous ...
    Indomethacin nanocrystals resulted in an increased solubility of 2.6 times, with a solubility of 17.2 ± 0.4 µg/mL. The highest increase was obtained with ...
  116. [116]
    Stability of Amorphous Pharmaceutical Solids: Crystal Growth ... - NIH
    We review recent progress toward understanding and enhancing the stability of amorphous pharmaceutical solids against crystallization.Fast Surface Crystal Growth · Surface Crystals Grow Upward... · Polymer Additives Can Have...<|control11|><|separator|>
  117. [117]
    Polymeric Amorphous Solid Dispersions: A Review of ...
    Jan 23, 2016 · Cui, Y. As a result, polymers prevent devitrification thereby preserving the viability (solubility and stability) of the amorphous state over ...
  118. [118]
    Role of polymers in the physical and chemical stability of amorphous ...
    Feb 1, 2022 · This paper shows that strong intermolecular interactions between CBZ and polymers contribute to the physical stability of the ASDs.Missing: devitrification | Show results with:devitrification
  119. [119]
    Stability and recrystallization of amorphous solid dispersions ...
    Mar 15, 2025 · The findings demonstrate that hot-melt extruded samples exhibit superior stability against recrystallization, whereas spray dried samples achieve higher ...
  120. [120]
    Development of Ternary Amorphous Solid Dispersions ...
    The objective of this study was to investigate the feasibility of a potential technology shift of spray-drying to HME in the process of the development of ASDs.Introduction · Experimental Section · Results · Discussion
  121. [121]
    Trends in amorphous solid dispersion drug products approved by ...
    Forty-eight (48) drug products (DPs) containing amorphous solid dispersions (ASDs) have been approved by the US Food and Drug Administration in the 12-year ...
  122. [122]
    Transformation of amorphous calcium phosphate to bone-like apatite
    Oct 9, 2018 · We demonstrate a detailed transformation mechanism of amorphous calcium phosphate spherical particles to apatite platelet-like crystals.
  123. [123]
    Amorphous calcium phosphate is a major component of the forming ...
    Sep 2, 2008 · Amorphous calcium phosphate is a major component of the forming fin bones of zebrafish: Indications for an amorphous precursor phase · Abstract.
  124. [124]
    Mapping amorphous calcium phosphate transformation into ...
    The possibility that calcium phosphate precursor phases are delivered and first deposited in bone has been debated for decades (1–3). In vitro precipitation of ...
  125. [125]
    Obsidian | Volcano World | Oregon State University
    Obsidian is natural glass that was originally molten magma associated with a volcano. This volcanic glass has an almost total absence of sizable mineral ...Missing: amorphous solid
  126. [126]
    Obsidian: A Pioneering Natural Resource for Green, Fire-Resistant ...
    Obsidian is characterized by its rigid, glassy composition with a brittle and amorphous texture, featuring razor-sharp edges due to the rapid cooling of silica ...
  127. [127]
    [PDF] Introduction - Rutgers University, GLASS |
    The cave-dwellers of even earlier times relied on chipped pieces of obsidian, a natural volcanic glass, for tools and weapons, i. e., scrapers, knives, axes, ...
  128. [128]
    [PDF] Tektites and their Origin - NASA Technical Reports Server (NTRS)
    Germanium and gallium content of tektites, impact glasses, meteorites and related materials. (Abstract). J. Geophys. Res., 65: 2482-2483. Cohen, A.J., 19. 6 ...
  129. [129]
    Quantitative Study of Porosity and Pore Features in Moldavites by ...
    Apr 24, 2014 · Tektites are today generally considered to be products of hypervelocity impacts on Earth; usually they experienced aerodynamic transport and can ...
  130. [130]
    A biomimetic peptide has no effect on the isotopic fractionation ...
    May 6, 2021 · Diatoms are encased in an outer cell wall, or frustule, made of hydrated amorphous silica ( n ). This is often referred to as opal or biogenic ...
  131. [131]
    Biogenic design of silicious architectures on Moso bamboo culm
    Jul 1, 2025 · Hydrated amorphous silica (SiO2•nH2O) is widely observed in a variety of living organisms, such as diatoms3,6,7,8,9,10,11, sponges ...
  132. [132]
    Humic Substances as a Versatile Intermediary - PMC - NIH
    Mar 23, 2023 · Hu, HA and FA occur naturally in the form of amorphous substances ([3]. Hu complexes are categorized as macro organic compounds due to their ...
  133. [133]
    Chemistry and potential mutagenicity of humic substances in waters ...
    Humic substances are amorphous organic macromolecules responsible for the hue of natural waters. They are also known to be precursors of mutagens formed on ...
  134. [134]
    Allophane isolated from a podsol developed on a non-vitric parent ...
    Mar 6, 1980 · The relationship between allophane, an X-ray amorphous, short-range ordered aluminosilicate gel, found in soils derived from volcanic ash ...
  135. [135]
    The viscosity of atmospherically relevant organic particles - Nature
    Mar 6, 2018 · This review explores our understanding of aerosol particle phase, particularly as identified by measurements of the viscosity of organic particles.
  136. [136]
    Sunlight can convert atmospheric aerosols into a glassy solid state ...
    Oct 21, 2022 · The increased viscosity likely leads to increased abundance of SOA in a glassy solid state that can potentially act as ice nucleation particles ...
  137. [137]
    Aging mechanisms in amorphous phase-change materials - Nature
    Jun 24, 2015 · This phenomenon is called physical aging and is usually attributed to structural relaxation towards the equilibrium state. Aging is ubiquitous ...
  138. [138]
    [PDF] Stretched Exponential Relaxation - arXiv
    Aug 2, 2018 · [9] (Friedrich. Kohlrausch also used the stretched exponential to describe mechanical stretching/creep relax- ations in the natural polymer silk ...
  139. [139]
    Predicting nonlinear physical aging of glasses from equilibrium ...
    The results in Fig. 3 demonstrate that nonlinear physical aging phenomena in the intermediate regime may be predicted from a knowledge of the linear limit of ...
  140. [140]
    Structural origins of Johari-Goldstein relaxation in a metallic glass
    Feb 3, 2014 · Johari-Goldstein or β relaxation, persisting down to glassy state from a supercooled liquid, is a universal phenomenon of glassy dynamics.
  141. [141]
    Impact of structural relaxation on mechanical properties of ...
    The degree of structural relaxation significantly affected the mechanical properties of copovidone as assessed by diametral compression tests, macroindentation ...
  142. [142]
  143. [143]
    Polyamorphism in water - PMC - NIH
    The transition pressure of ~0.6 GPa was much lower than that of ~1 GPa of ice Ih, which hinted that LDA was truly amorphous. Then, the halo pattern of LDA ...
  144. [144]
  145. [145]
  146. [146]
  147. [147]
  148. [148]