Fact-checked by Grok 2 weeks ago

Fracture

A fracture is the separation of an object or material into two or more pieces under the action of applied , often appearing as a or complete break. In and , the study of fractures is central to , a field that analyzes , , and to predict and prevent structural breakdowns in components like bridges, , and machinery. Fractures are broadly classified into two types: brittle, where occurs with little or no plastic deformation, leading to sudden and catastrophic s; and ductile, involving significant plastic deformation before separation, often resulting in a more gradual mode. Understanding these mechanisms is crucial for designing safer materials and assessing risks in high-stress environments, as uncontrolled fractures have contributed to numerous .

Fundamentals

Definition and Scope

Fracture is the irreversible separation of a solid material into two or more parts when subjected to applied that exceeds its capacity to withstand loading, resulting in breaking or cracking. This process fundamentally differs from deformation, which involves reversible straining, or yielding, which allows permanent shape change without complete separation. The theoretical foundation of fracture mechanics originated with A.A. Griffith's 1921 work on brittle fracture in , where he analyzed the propagation of pre-existing cracks through an energy balance approach, explaining why brittle materials fail at stresses far below their theoretical strength. Griffith's criterion for the onset of brittle fracture provides a quantitative relation, expressed as: \sigma_f = \sqrt{\frac{2E\gamma}{\pi a}} where \sigma_f denotes the fracture stress, E is the , \gamma represents the surface energy required to create new surfaces, and a is the half-length of an internal (or the full length for an ). This equation highlights the critical role of flaw size in determining material strength, shifting focus from uniform material properties to defect-controlled . The scope of fracture studies in and broadly applies to solid materials such as metals, ceramics, and polymers, where modes range from rapid brittle separation to more gradual ductile processes. Understanding fracture is essential for designing reliable components, as it informs strategies to mitigate risks in applications from structural parts to biomedical implants, distinguishing from controlled deformation.

Basic Principles of Fracture Mechanics

Linear elastic fracture mechanics (LEFM) forms the cornerstone of modern fracture analysis, focusing on the behavior of cracks in materials that remain predominantly elastic. Originating from A.A. Griffith's pioneering work on brittle fracture in 1921, LEFM was formalized by G.R. Irwin in the 1950s to address the stress concentration and propagation criteria for cracks under linear elastic conditions. LEFM assumes small-scale yielding, meaning the region of plastic deformation at the crack tip is much smaller than both the crack length and the overall specimen dimensions, allowing linear elasticity to govern the far-field response while capturing the intense local stresses. Additionally, the material is assumed to behave elastically away from the crack tip, with quasi-static loading and no significant time-dependent effects. These assumptions enable predictive models for crack stability and growth based on continuum mechanics. Central to LEFM is the stress intensity factor K, which quantifies the magnitude of the three-dimensional stress field near the crack tip and serves as a fracture criterion. Irwin defined three primary modes of crack loading: mode I for tensile opening normal to the crack plane, mode II for in-plane sliding or , and mode III for out-of-plane tearing or anti-plane . For an infinite plate containing a central through-crack of length $2a subjected to uniform remote tensile \sigma perpendicular to the crack, the mode I is expressed as K_I = \sigma \sqrt{\pi a}. This formulation highlights how K integrates the effects of applied stress, crack length, and geometry, with crack propagation initiating when K exceeds a material-specific critical value K_c. The near-tip stresses follow a singular form \sigma_{ij} \sim K / \sqrt{2\pi r}, where r is the radial distance from the tip, underscoring the infinite stress concentration in ideal elastic theory. Complementing the stress-based approach, LEFM employs an energy balance criterion rooted in Griffith's theory, where stable crack growth requires the release of elastic strain to overcome the surface of new crack faces. The release rate G, defined as the decrease in per unit crack advance, governs fracture when G \geq G_c, the critical release rate. Irwin linked this to the stress intensity factor through G = \frac{K^2}{E'}, where E' is the effective modulus: E' = E under and E' = E / (1 - \nu^2) under plane strain, with E as and \nu as . This equivalence bridges stress and perspectives, enabling unified criteria for brittle fracture prediction. The idealized singularity at the crack tip is moderated in real materials by localized , forming a small zone where stresses are capped by the yield strength \sigma_y. Irwin estimated the plane-stress plastic zone radius along the crack plane as r_p \approx \frac{1}{2\pi} \left( \frac{K}{\sigma_y} \right)^2, derived by setting the elastic \sigma_{yy} stress equal to \sigma_y and solving for the distance r where yielding begins. Under plane strain, the zone is roughly one-third smaller due to triaxiality constraints. This size correction validates LEFM applicability when r_p \ll a, ensuring the plastic enclave does not perturb the elastic K-dominated . For ductile materials exhibiting extensive , where r_p approaches or exceeds structural dimensions, LEFM's small-scale yielding assumption fails, prompting a shift to elastic-plastic (EPFM). EPFM extends LEFM principles to nonlinear regimes using path-independent integrals like the , introduced by J.R. Rice in 1968, which generalizes [G](/page/G) for incremental plasticity and characterizes crack-tip driving force under large deformation. This transition is essential for metals and alloys where yielding precedes unstable fracture.

Material Response to Stress

Fracture Strength

Fracture strength, also known as breaking strength, refers to the maximum stress a can endure immediately prior to fracturing under tensile loading, marking the point of complete failure. This differs from yield strength, which indicates the onset of permanent deformation without fracture, allowing materials—particularly ductile ones—to sustain loads beyond yielding before breaking. In brittle , fracture strength often coincides with , as failure occurs abruptly without significant , whereas in ductile materials, it follows necking after reaching the ultimate stress peak. Microstructural features profoundly influence fracture strength, with grain size, defects, and temperature playing key roles. The Hall-Petch relation empirically captures the strengthening effect of finer grains, where fracture strength \sigma_f increases inversely with the of grain diameter d: \sigma_f = \sigma_0 + k d^{-1/2}, with \sigma_0 as the friction stress and k as the strengthening coefficient; this arises from increased barriers to motion and propagation. Defects such as voids, inclusions, and microcracks act as stress concentrators, drastically lowering strength by facilitating premature initiation. Elevated temperatures generally reduce fracture strength by enhancing atomic mobility, promoting glide, and accelerating diffusion-mediated processes like , though specific effects vary by material class. For brittle materials exhibiting stochastic failure due to inherent flaw distributions, fracture strength follows a statistical description via the Weibull distribution, which models the probability of failure P_f under uniform stress \sigma over volume V:
P_f = 1 - \exp\left[-\left(\frac{V}{V_0}\right)\left(\frac{\sigma}{\sigma_0}\right)^m\right],
where V_0 is a reference volume, \sigma_0 a characteristic strength, and m the Weibull modulus reflecting flaw variability (higher m indicates more consistent strength). This "weakest-link" approach accounts for size effects, where larger volumes increase failure likelihood at lower stresses./06%3A_Yield_and_Fracture/6.03%3A_Statistics_of_Fracture)
Environmental factors like and significantly degrade fracture strength by introducing surface degradation and internal stresses. , through pitting or uniform attack, creates localized stress raisers that reduce effective cross-section and initiate cracks, often lowering strength by 20-50% depending on exposure duration and severity. , prevalent in high-strength steels, diffuses atomic into the lattice, promoting brittle and reducing ; it can diminish fracture strength by up to 50% or more in susceptible alloys by facilitating hydrogen-enhanced decohesion or localized . The disparity between theoretical and actual fracture strength underscores the role of imperfections. In an ideal, flaw-free , theoretical strength approximates E/10 (where E is the ), derived from the needed to break atomic bonds uniformly. However, real materials achieve only about E/1000 due to preexisting flaws, as explained by Griffith's criterion, which posits that cracks propagate when the intensity overcomes , yielding strengths orders of magnitude below the theoretical limit.

Fracture Toughness and Energy Absorption

quantifies a material's resistance to the propagation of a under applied , particularly in the linear elastic (LEFM) regime where plastic deformation is minimal. The critical , K_{Ic}, represents this resistance for mode I (opening mode) loading under plane-strain conditions, defined as the stress intensity at which a extends unstably. Its units are MPa\sqrt{\mathrm{m}}, reflecting the combination of and crack length scales that govern crack-tip fields. In elastic-plastic fracture mechanics (EPFM), where plays a significant role, the serves as a key parameter to characterize crack driving force and energy release. Introduced by in 1968, the is a path-independent contour integral that measures the energy available for crack advance per unit crack extension, applicable to nonlinear elastic or elastic-plastic materials. It is mathematically expressed as J = \int_{\Gamma} \left( \gamma \, ds + \mathbf{T} \frac{\partial \mathbf{u}}{\partial a} \cdot \mathbf{n} \, dl \right), where \Gamma is a surrounding the crack tip, \gamma is the strain density, \mathbf{T} is the traction , \mathbf{u} is the displacement , a is the length, and \mathbf{n} is the unit outward normal. For steady-state crack growth, J equals the rate of dissipation per unit crack advance, providing a fracture when compared to a material's J_{Ic}. Plasticity contributes substantially to energy absorption during fracture, particularly in ductile materials, by enabling mechanisms such as void , , and coalescence ahead of the crack tip. In ductile fracture, voids form at inclusions or defects and expand under triaxial stress, eventually coalescing to create a fracture surface that dissipates through extensive deformation. bands, localized regions of intense plastic shear, further enhance energy absorption by concentrating deformation and facilitating crack path deviation, often leading to dimpled fracture surfaces. These processes allow ductile materials to absorb significantly more —up to orders of magnitude higher—compared to brittle ones before . Fracture toughness exhibits strong dependence on temperature and loading rate, influencing the balance between ductile and brittle behavior. In body-centered cubic (BCC) metals like steels, the ductile-to-brittle transition temperature (DBTT) marks the point where toughness drops sharply as dislocation mobility decreases at lower temperatures, shifting failure from energy-absorbing ductile modes to low-toughness . For ferritic steels, the DBTT typically ranges from -50°C to 50°C depending on composition and microstructure, with higher strain rates elevating it by restricting plastic flow. Standardized measurement of K_{Ic} in the LEFM follows ASTM E399, which specifies single-edge-notched compact tension specimens to ensure plane-strain conditions and valid values. The requires precracking, monotonic loading to instability, and validation checks such as a minimum thickness to suppress plastic zone effects, ensuring the measured K_{Ic} reflects intrinsic material resistance rather than geometric influences.

Types of Fracture

Brittle Fracture

Brittle fracture occurs with negligible deformation, leading to abrupt and often as cracks propagate rapidly through the material. In such failures, the fracture surface exhibits along specific crystallographic planes, typically following transgranular paths that minimize energy absorption during propagation. This mode is common in materials like body-centered cubic (BCC) metals, ceramics, and , where atomic bonding favors separation over dislocation-mediated deformation. Key causes of brittle fracture include exposure to low temperatures, high strain rates, and impurities in BCC metals that elevate the ductile-to-brittle transition temperature (DBTT). For instance, elements like segregate to grain boundaries, promoting intergranular weakening and increasing the DBTT, which shifts the material toward brittle behavior under service conditions. A historical example is the brittle failures of ships during , where mild hulls fractured suddenly in cold North Atlantic waters due to its inherent notch sensitivity and the combined effects of low temperatures and welding-induced stress concentrations. Crack propagation in brittle fracture proceeds at speeds approaching the surface wave speed, roughly 0.9 times the shear wave speed (c_s), enabling near-instantaneous failure across large sections. Representative examples include the shattering of under tensile loading, where mirror-like cleavage facets form on the fracture surface, and the cracking of ceramics like alumina under similar stresses, often initiating from surface flaws. Prevention of brittle fracture focuses on metallurgical and measures, such as alloying BCC steels with to lower the DBTT and enhance low-temperature , or components to minimize tensile stresses at potential crack sites through rounded features and crack arrestors. Unlike ductile fracture, which allows gradual energy dissipation via plastic flow, brittle fracture provides little warning before complete separation.

Ductile Fracture

Ductile fracture occurs when a undergoes substantial deformation before final separation, allowing for energy dissipation through mechanisms that prevent sudden failure. This mode of failure is prevalent in metals and alloys capable of extensive yielding, where the fracture surface appears rough and fibrous due to the stretching and tearing of the . Unlike brittle fracture, ductile failure involves a gradual process that provides warning through visible deformation, making it less catastrophic in many applications. The progression of ductile fracture typically unfolds in distinct stages, beginning with necking, where localized reduction in cross-sectional area occurs under tensile loading due to . This is followed by , often initiated at inclusions, second-phase particles, or microstructural defects that act as stress concentrators during plastic straining. Subsequent void growth expands these cavities as surrounding material deforms plastically, and finally, coalescence links adjacent voids, forming a continuous that propagates to complete separation. These stages are well-documented in experimental observations of metals under monotonic loading. Central to understanding ductile fracture is the microvoid coalescence model, which describes how voids nucleate, grow, and merge to produce characteristic fracture surface features. As voids enlarge and impinge, they create equiaxed or elongated dimples on the fracture surface, reflecting the flow around the cavities. In standard tensile tests of cylindrical specimens, this mechanism yields the iconic cup-and-cone morphology: a central "cup" region with radial dimples from axisymmetric void growth, surrounded by a "" of lips where oblique fracture occurs due to -dominated coalescence. This model, supported by scanning electron of fracture surfaces, highlights the role of triaxial states in accelerating void linkage. Several factors influence the propensity for ductile fracture, particularly in face-centered cubic (FCC) metals such as aluminum, which exhibit high owing to their multiple slip systems that facilitate glide. Elevated temperatures further promote by increasing atomic mobility, reducing the critical stress for motion, and delaying the onset of void coalescence. Additionally, strain hardening enhances overall ; as multiply and tangle, the material's rises, enabling greater uniform elongation before necking and fracture initiation. These effects are evident in aluminum alloys, where higher strain-hardening exponents correlate with improved resistance to localized failure. Energy dissipation during ductile fracture primarily occurs through irreversible plastic deformation processes, including the generation and motion of dislocations that accommodate . Work hardening contributes significantly by storing energy in the lattice via dislocation interactions, while frictional losses from dislocation glide further absorb applied work, allowing the material to deform extensively without immediate rupture. This dissipation mechanism underpins the higher fracture energy of ductile materials compared to brittle ones. Practical examples of ductile fracture include the tearing of metal components under overload, such as in structural steels where excessive tensile forces cause necking and dimpled rupture. In welded assemblies, overload can lead to ductile failure along the weld toe or , manifesting as tearing with significant flow before separation, as observed in fillet welds under extreme loading.

Fracture Characteristics

Macroscopic Observations

Macroscopic observations of fracture surfaces provide critical insights into the failure mode and loading conditions without requiring magnification, revealing patterns that distinguish between brittle, ductile, and -induced fractures. In fractures, beach appear as concentric ridges or lines on the surface, originating from the crack initiation site and indicating progressive under cyclic loading. These are often visible as semi-elliptical patterns that fan out from multiple origins in cases of multiple crack starts, aiding in identifying the origin. Brittle fractures in metals exhibit river patterns, which manifest as branching, feather-like lines radiating from the crack origin on the fracture surface, characteristic of failure along crystallographic planes. These patterns form due to the rapid propagation of cracks in low-ductility materials, creating a textured appearance that contrasts with smoother ductile surfaces. In ceramics, fracture surfaces display distinct zones including and regions; the zone appears as a hazy transition from the initial smooth mirror area, while zones show coarse, irregular ridges indicating higher crack velocities near the point of branching. These features correlate with increasing crack speed, with formation signaling velocities approaching 0.5 to 0.8 times the speed in the material. Ductile fractures are marked by evident plastic deformation, such as necking—a localized in cross-sectional area—and lips, which are slanted, fibrous edges at the fracture resulting from shear-dominated final separation. These signs produce a rough, dimpled surface with significant gross deformation, contrasting sharply with the flat, shiny appearance of brittle fractures where minimal leads to perpendicular, granular planes with high reflectivity. In and brittle ceramics, the classic mirror-mist-hackle sequence delineates the fracture progression: the mirror is a smooth, featureless zone near the reflecting slow initial growth, transitioning to the misty, diffuse region, and culminating in the rough, ridged area before macroscopic branching occurs. This sequence allows estimation of fracture based on mirror measurements. Post-fracture often involves matching complementary fracture surfaces from separated components to reconstruct the failure , using techniques like Fracture Surface (FRASTA) to align 3D topographies and trace paths. These macroscopic matches reveal the of propagation and loading history, linking surface features to incident specifics without delving into microscopic .

Microscopic Mechanisms

At the scale, fracture and in crystalline materials are governed by the behavior of at crack tips, where concentrations drive either emission or atomic bond breaking. pile-ups form when mobile dislocations accumulate under applied , creating intense local stresses that can either emit further dislocations or initiate . The driving dislocation motion in such configurations is described by the Peach-Koehler force, given by \mathbf{F} = (\boldsymbol{\sigma} \cdot \mathbf{b}) \times \boldsymbol{\xi}, where \boldsymbol{\sigma} is the tensor, \mathbf{b} is the , and \boldsymbol{\xi} is the line ; this balances and near the crack tip to determine . Seminal analyses of pile-ups, such as those modeling linear arrays against barriers like crack tips, show that the ahead of the pile-up scales inversely with the of distance, amplifying the for . Atomic-scale models further elucidate the competition between dislocation emission and cleavage decohesion. The Rice-Thomson criterion posits that brittle fracture occurs if the energy barrier for spontaneous emission from an atomically sharp crack exceeds thermal activation at relevant temperatures, while ductile behavior prevails if emission blunts the crack readily; this is quantified by the ratio \mu b / \gamma, where \mu is the , b the , and \gamma the surface , with values below approximately 10 favoring emission in metals like . In face-centered cubic crystals, wide dislocation cores and favorable slip plane orientations relative to the crack facilitate emission, whereas narrow cores in covalent materials like promote . This criterion highlights that even in nominally ductile materials, high stress intensity can suppress emission, leading to if the critical emission distance exceeds the atomic core size. Grain boundaries significantly influence fracture paths by altering local dislocation dynamics and cohesion. Intergranular fracture predominates when impurities segregate to boundaries, reducing cohesive strength through charge transfer mechanisms that weaken metal-metal bonds, as observed in nickel with sulfur or iron with phosphorus; this shifts failure from transgranular cleavage across grains to brittle separation along boundaries. Segregation-induced embrittlement is exacerbated by boundary misorientation and second-phase particles, which impede dislocation transmission across boundaries, increasing the propensity for intergranular paths over transgranular ones that involve dislocation-mediated plasticity within grains. In contrast, clean boundaries in pure metals favor transgranular fracture due to higher intrinsic cohesion. At the nanoscale, crack tip blunting in ductile materials arises from dislocation emission and subsequent plastic shearing, which rounds the sharp tip and dissipates energy. In nanocrystalline platinum, atomistic simulations reveal alternating sequences of dislocation nucleation, glide, and annihilation at the tip, leading to repeated blunting that arrests propagation and enhances toughness; this process is driven by shear stresses exceeding lattice resistance, with blunting radii scaling with grain size. Such mechanisms underscore why ultrafine-grained metals exhibit improved ductility, as confined dislocations promote emission over cleavage. Deformation twins and phase transformations further modulate fracture paths, particularly in steels, by redirecting crack and absorbing energy. In body-centered cubic steels, twinning induced by high strain rates creates barriers that deflect cracks, increasing and , as twins act as planar obstacles to motion similar to low-angle boundaries. Martensitic phase transformations in advanced high-strength steels play a : they initially blunt cracks via volume expansion during austenite-to-martensite , shielding the , but can later promote under cyclic loading by generating transformation-induced stresses that exceed local . These effects are prominent in transformation-induced plasticity steels, where controlled stability of the phase optimizes both twinning and transformation for enhanced fracture resistance.

Testing Methods

Experimental Techniques

Experimental techniques for studying fractures in materials involve controlled loading of specimens to induce while deformation and propagation. These methods allow researchers to replicate fracture conditions in settings, providing data on material behavior under various stress states. Key approaches include protocols that apply tensile, bending, or impact loads, often combined with real-time observation tools to capture dynamic processes. Sample preparation is crucial for ensuring reproducible and controlled crack initiation. Notched specimens, such as single-edge-notched bend (SENB) or compact geometries, are machined with precise V-shaped or U-shaped es to concentrate and promote growth from a defined . precracking is commonly applied to sharpen the notch into a natural , minimizing artificial effects and enabling accurate of real-world defects. This preparation follows guidelines in standards like those from ASTM, ensuring consistency across tests. Tensile testing applies uniaxial loads to elongate specimens until fracture, generating stress-strain curves that reveal yield strength, ultimate tensile strength, and ductility. The ASTM E8/E8M standard specifies procedures for metallic materials, including specimen dimensions (e.g., round or flat geometries) and loading rates to achieve quasi-static conditions. During the test, extensometers measure strain, and failure typically occurs via necking followed by ductile or brittle rupture, depending on the material. This method is widely used to assess fracture initiation under monotonic loading. Bend tests evaluate fracture in beam-like specimens by applying transverse loads, measuring load-deflection responses to determine and energy absorption. In three-point , a central load is applied over two supports, creating maximum at the midpoint, while four-point distributes the load over a wider region for more uniform fields. ASTM standards such as E290 flexural testing, with notched beams often used to study crack propagation under moments. These tests are particularly useful for brittle materials where may be impractical due to gripping issues. Impact testing assesses fracture behavior at high strain rates using pendulum devices to simulate sudden loading. The Charpy test involves a swinging hammer striking a notched bar supported at both ends, measuring the energy absorbed during fracture as the pendulum's height difference. Similarly, the test uses a setup for the same purpose. Governed by ASTM E23, these methods quantify transitions, such as in steels where drops at low temperatures. Results are reported in joules, providing insights into dynamic fracture resistance. In-situ observation techniques enable direct visualization of fracture mechanisms during loading. Scanning electron microscopy (SEM) integrated with loading stages captures microcrack initiation, propagation, and coalescence in real time, often at magnifications up to 10,000x. (AE) monitoring complements this by detecting high-frequency waves from crack growth or motion using piezoelectric sensors. AE signals are analyzed for , duration, and count to correlate events with fracture stages, offering non-destructive insights into damage accumulation. These methods, when combined, provide a comprehensive view of evolving microstructures without interrupting the test.

Fracture Toughness Measurement

Fracture toughness measurement involves standardized testing protocols to quantify a material's resistance to crack propagation under controlled conditions, primarily focusing on the critical stress intensity factor K_{Ic} for linear-elastic fracture mechanics (LEFM) and related parameters for elastic-plastic fracture mechanics (EPFM). These methods ensure reproducibility and validity by specifying specimen geometries, loading configurations, and acceptance criteria, allowing comparison across materials and applications. The primary standard for metallic materials is ASTM E399, which outlines procedures for determining plane-strain fracture toughness using fatigue-precracked specimens. In ASTM E399, testing typically employs single-edge notched bend (SENB) or compact tension () specimens, where a sharp is introduced via fatigue precracking to simulate realistic flaw conditions. The specimen is loaded in three-point bending for SENB or tensile loading for , with load and displacement recorded to calculate the K_Q at the point of initiation, defined as 1.99% apparent compliance increase or a specific load line displacement. If validity criteria are met, K_Q is accepted as K_{Ic}. Key validity requirements include ensuring minimal growth stability, where the maximum load P_{\max} satisfies P_{\max} / P_Q \leq 1.10, indicating that the test remains within LEFM assumptions without excessive . Additionally, plane-strain conditions must prevail to obtain a size-independent property, requiring specimen thickness B, length a, and remaining W - a to each exceed $2.5 (K_{Ic} / \sigma_y)^2, where \sigma_y is the yield strength; this ensures the plastic zone at the tip is constrained, minimizing triaxiality effects. For materials exhibiting significant where LEFM assumptions fail, EPFM methods are employed to characterize growth resistance through the , as detailed in ASTM E1820. The J_R curve plots J (energy release rate) against extension \Delta a, providing a resistance curve that captures rising with growth due to plastic work. In the multiple-specimen , several identically prepared specimens (often or SENB) are loaded to predetermined extensions, unloaded, and the initial and final lengths measured post-fracture via optical or methods to construct the J_R curve; this approach avoids real-time monitoring challenges in ductile materials. The initiation J_{Ic} is determined at a extension of 0.2 mm or using an offset from blunting line, offering a measure of growth resistance beyond plane-strain limits. Dynamic fracture toughness assessment is critical for high-strain-rate applications, such as impact or ballistic loading, where rate sensitivity alters crack propagation. The (SHPB), also known as the Kolsky bar, is a widely adopted technique for these measurements, involving a striker bar generating a compressive transmitted through incident and transmitter bars to dynamically load a notched specimen at rates exceeding $10^3 s^{-1}. Configurations like modified three-point bend or semi-circular bend specimens allow determination of dynamic K_{Id} by analyzing wave propagation, load history, and post-test to quantify crack speed and toughness, often revealing rate-dependent increases in for metals and polymers. Validity requires stress equilibrium and minimal in the specimen, with results interpreted using one-dimensional wave theory. Interpreting fracture toughness data necessitates accounting for size effects and factors to ensure applicability to structural components. Smaller specimens may yield higher apparent due to plane-stress conditions, where the plastic zone extends unconstrained, inflating K values; thus, to larger sizes via -adjusted models is essential for conservative . factors, such as the T-stress or Q-parameter, quantify triaxiality at the crack tip, with loss of in shallow-notched or large-scale tests leading to elevated by 20-50% compared to deeply notched specimens; corrections using these factors enable transferability from test to component . For instance, in ferritic steels, statistical size effects from weakest-link models further modulate lower-bound in larger volumes.

Fracture in Specific Materials

Ceramics and Inorganic Glasses

Ceramics and inorganic glasses exhibit inherently brittle fracture behavior, characterized by low fracture toughness values typically ranging from 1 to 5 MPa√m, which limits their ability to absorb energy before catastrophic failure. This brittleness arises from strong ionic and covalent bonding that restricts plastic deformation, making these materials highly sensitive to preexisting flaws. Flaw sensitivity in ceramics is quantitatively described by Weibull statistics, which model the probabilistic nature of fracture based on the weakest-link theory, where the failure probability depends on the volume or surface distribution of critical defects. The Weibull modulus, often between 5 and 15 for polycrystalline ceramics, indicates the scatter in strength data, with lower values signifying greater variability due to flaw populations. A key challenge in ceramics is their vulnerability to , where rapid temperature changes induce tensile stresses that can initiate s from surface flaws. resistance is assessed using the R = \frac{\sigma_f (1 - \nu)}{E \alpha}, where \sigma_f is , \nu is , E is , and \alpha is the coefficient of ; higher R values predict better resistance to initiation. This parameter highlights trade-offs, such as how low \alpha in materials like enhances resistance despite moderate strength. defects, including pores and inclusions, serve as primary nuclei in these materials, as incomplete densification during processing leaves voids or foreign particles that act as stress concentrators. Pores, often equiaxed and up to 20 μm in size, promote , while inclusions like in can trigger spontaneous under residual stresses. To mitigate brittleness, toughening methods exploit extrinsic mechanisms, such as phase in zirconia ceramics, where stress-induced tetragonal-to-monoclinic around tips generates compressive stresses that shield the front and increase effective toughness to 5-10 MPa√m. This seminal mechanism, first demonstrated in partially stabilized zirconia, expands transformation zones up to 10-20 μm wide, dissipating energy through volume dilation of about 4%. bridging by elongated grains or particles further enhances resistance by applying closure tractions behind the tip, as seen in where grain pullout contributes up to 50% of the toughening increment. In practical applications, such as ceramic tiles and glass windows, fracture often results from slow growth under , where water-assisted bond rupture at tips follows a power-law relation, leading to subcritical extension over time and eventual under service loads like or thermal cycling. For instance, panels exhibit delayed fracture from environmental exposure, with velocities accelerating from 10^{-10} to 10^{-3} m/s as approaches the critical value.

Fiber-Reinforced Composites

Fiber-reinforced composites (FRCs) are engineered materials consisting of high-strength fibers embedded in a , designed to enhance and strength compared to monolithic matrices. In these systems, fracture arises from interactions between the fibers and matrix under mechanical loading, leading to progressive damage rather than sudden brittle . The anisotropic nature of FRCs, particularly in laminate or bundle configurations, results in fracture behaviors influenced by fiber , properties, and loading . Key failure modes in FRCs include fiber breakage, where individual fibers fracture under excessive tensile or ; matrix cracking, initiating as microcracks in the or that propagate and coalesce; , the separation of layered plies due to interlaminar stresses; and fiber pull-out, where fibers debond from the and slide out, dissipating energy. These modes often occur sequentially, with matrix cracking preceding fiber-dominated failures in tension-loaded unidirectional composites. For instance, in carbon fiber-reinforced (CFRP) laminates, is a major contributor to the total fracture energy in mode I loading scenarios. Bundle theory models fracture in FRCs by treating parallel fiber arrays as load-sharing systems, where the failure of weaker fibers redistributes to survivors, leading to a characteristic global load-displacement curve with an initial linear rise followed by nonlinear softening due to cascading . In equal load-sharing variants, the bundle's overall strength is determined by the statistical of fiber thresholds, providing a simple framework for predicting composite tensile . Local load-sharing extensions account for stress concentrations near broken fibers, accelerating localization in real composites. General fiber bundles serve as foundational models for these behaviors, illustrating avalanche-like . Toughening in FRCs primarily occurs through mechanisms such as fiber bridging, where intact fibers span faces and transfer load across the fracture plane, increasing resistance to crack growth; and crack deflection at fiber-matrix interfaces, which forces cracks to deviate from straight paths, extending the fracture surface and absorbing energy. These extrinsic toughening effects can elevate the of brittle matrices by factors of 10 or more, as seen in ceramic-matrix composites where bridging contributes significantly to the total work of fracture. Weak interfaces promote debonding and pull-out, further enhancing energy dissipation without premature fiber fracture. Micromechanically, the estimates composite strength as a volume-weighted of and contributions, assuming perfect load ; however, in practice, fracture is often governed by weak interfaces that initiate debonding and limit overall performance. This discrepancy highlights the role of interface shear strength, where values below 10 can shift failure from breakage to matrix-dominated modes, reducing effective . In applications, carbon fiber-reinforced composites exemplify these fracture characteristics, with components like fuselages experiencing tow bundle failures under tension due to bundle pull-out, where clustered fibers debond collectively, leading to sudden load drops. studies on CFRP underscore the need for optimized interfaces to mitigate such risks in high-stress environments.

Advanced Applications and Analysis

Case Studies in Disasters

During , the mass production of over 2,700 ships for the U.S. Merchant Marine resulted in numerous brittle fractures in their welded steel hulls, particularly during cold weather operations. The steel alloy contained high levels of impurities like and , which elevated the ductile-to-brittle transition temperature and rendered the material prone to sudden cracking under tensile stresses from waves or structural loads. processes further contributed by creating brittle heat-affected zones and residual stresses that initiated cracks at discontinuities such as holes or deck edges. Historians have documented 19 instances of ships splitting in two without warning, leading to the loss of 12 vessels and the deaths of at least 13 crew members in related incidents. In 1954, two 1 jetliners experienced explosive decompression and mid-air disintegration due to propagation around square passenger windows in the pressurized . The rectangular design of the windows introduced severe stress concentrations at their corners, accelerating growth during repeated pressurization-depressurization cycles far below the anticipated life of 16,000 flights. On January 10, broke apart at 27,000 feet near , , killing all 35 occupants; on April 8, Flight 201 disintegrated at 35,000 feet near , , resulting in 21 fatalities. These accidents prompted the indefinite grounding of the global fleet and redesigns incorporating rounded windows to distribute stresses more evenly. The Hyatt Regency Hotel walkway collapse on July 17, 1981, in , demonstrated how a design modification in steel rod connections could lead to ductile overload and structural failure under unexpected loads. The original engineering plans specified continuous hanger rods passing through both the second- and fourth-floor walkway beams for support from the atrium ceiling, but fabricators proposed—and engineers approved—a change to separate rods for each walkway, which halved the connection capacity and imposed twice the intended shear load on the upper-level brackets. During a crowded , the weakened fourth-floor connections sheared off, causing both walkways to plummet and resulting in 114 deaths and over 200 injuries—the deadliest non-terrorism structural failure in U.S. history at the time. In a contemporary , the September 9, 2010, rupture of a 30-inch transmission in , highlighted fracture risks from -induced combined with manufacturing defects. Operated by Pacific Gas and Electric, the pipeline segment featured a poor-quality longitudinal seam weld that allowed internal and external damage to propagate a longitudinal over decades, exacerbated by inadequate pressure testing and record-keeping. The failure released 47 million standard cubic feet of gas, which ignited and created a massive , destroying 38 homes, killing 8 people, hospitalizing 51 others, and causing over $220 million in property damage. These disasters collectively emphasize the critical need for fracture-safe , which requires engineers to select materials with low nil-ductility transition temperatures, perform flaw-tolerant analyses accounting for levels and potential sizes, and ensure structures operate above brittle transition ranges to or prevent fracture initiation. Equally vital is the routine application of non-destructive testing (NDT) techniques, such as ultrasonic and methods, to detect and size subsurface cracks early, enabling assessments that predict growth rates and maintain safety factors—preventing failures like those in pipelines or by verifying weld integrity without compromising components.

Computational Fracture Mechanics

Computational fracture mechanics encompasses numerical techniques to simulate crack initiation, propagation, and interaction in materials, enabling predictions of fracture behavior under various loading conditions without relying solely on experimental trials. These methods bridge theoretical fracture principles, such as stress intensity factors, with practical engineering simulations by discretizing the domain and solving governing equations computationally. Key approaches include enrichment-based finite element methods, boundary integral formulations, diffuse interface models, and hierarchical multiscale couplings, each tailored to handle the discontinuities and nonlinearities inherent in fracture processes. The (FEM) forms a cornerstone of computational fracture simulations, particularly through extensions that accommodate arbitrary crack paths without mesh modification. The (XFEM) enriches standard finite element approximations with discontinuous functions, such as the Heaviside step for crack interiors and asymptotic near-tip fields, allowing crack tracking independent of the underlying mesh. This avoids costly remeshing, making it suitable for complex geometries and evolving cracks. Complementary to XFEM, cohesive zone models (CZMs) within FEM frameworks represent interface failure by inserting zero-thickness elements along potential fracture surfaces, governed by traction-separation laws that capture progressive damage from initial elastic response to complete decohesion. These models effectively simulate and ductile fracture by regularizing the singularity at crack tips. The (BEM) offers an alternative for linear elastic fracture problems, formulating the problem in terms of boundary integrals rather than , thereby reducing the dimensionality from three to two (or two to one) and minimizing computational overhead for infinite or semi-infinite . In fracture applications, hypersingular and displacement-based integral equations handle crack-face tractions and displacements, enabling efficient analysis of multiple and computations. Dual BEM formulations address non-uniqueness issues in interior points, enhancing accuracy for and crack problems. Phase-field models provide a diffuse representation of cracks by introducing a continuous variable φ (where φ=0 denotes intact and φ=1 fully broken), embedded within a variational functional F that minimizes both elastic strain and fracture . Crack evolution follows an Allen-Cahn-type equation, ∂φ/∂t ∝ -δF/δφ, where the drives irreversible , regularizing sharp over a small length scale. This approach naturally handles crack branching, , and complex topologies without explicit tracking, though it requires careful calibration of the regularization parameter to match Griffith's criterion. Multiscale approaches couple atomistic simulations, such as , with models to resolve fine-scale phenomena at crack tips, where assumptions break down due to discreteness and nonlinearity. Concurrent methods, like the bridging domain approach, overlap atomistic and regions, partitioning the to ensure and seamless transition of displacements and forces across scales. This enables accurate prediction of emission and lattice trapping effects influencing speeds. Recent advances (as of 2025) incorporate physics-based techniques, such as deep neural networks and , to enhance multi-scale simulations of fracture processes. These methods improve flexibility and data-driven predictions of crack paths and material behavior under complex loading, reducing computational costs for large-scale problems. Validation of these computational models often involves benchmarking against standardized experiments, such as double cantilever beam (DCB) delamination tests, which measure mode I in composites. For instance, XFEM and CZM simulations of DCB specimens accurately reproduce load-displacement curves and crack growth rates observed in carbon-epoxy laminates, with errors below 5% when interface properties are calibrated from optical measurements. Similarly, phase-field predictions align with DCB data by adjusting the crack length scale to experimental limits, confirming robustness for quasi-static .

References

  1. [1]
    Broken bone: MedlinePlus Medical Encyclopedia
    ### Summary of Broken Bones (MedlinePlus)
  2. [2]
    Current Options for Determining Fracture Union - PMC - NIH
    Sep 14, 2014 · Introduction. There are about 6 million fractures in the United States annually and 5–10% of these fractures proceed to nonunion [1]. The risk ...4. Measures Of Healing · 4.1. Imaging Measures · 4.1. 1. Radiographic Union...
  3. [3]
    Bone Fractures: Types, Symptoms & Treatment - Cleveland Clinic
    A bone fracture is the medical definition for a broken bone. Fractures are usually caused by traumas like falls, car accidents or sports injuries.Greenstick Fractures · Oblique Fracture · Spiral Fracture · Buckle Fracture
  4. [4]
    The Frequency of Bone Disease - Bone Health and Osteoporosis
    Each year an estimated 1.5 million individuals suffer a fracture due to bone disease. The risk of a fracture increases with age and is greatest in women.Fractures Caused by Bone... · Osteoporosis · Other Metabolic Bone Diseases
  5. [5]
    Fracture Healing Overview - StatPearls - NCBI Bookshelf
    Apr 8, 2023 · [1] A fracture is a breach in the structural continuity of the bone cortex, with a degree of injury to the surrounding soft tissues.
  6. [6]
    [PDF] 1 CHAPTER 11 FRACTURE OF MATERIALS 11.1 Brittle vs. Ductile ...
    Fracture involves the forced separation of a material into two or more parts. Brittle Fracture involves fracture without any appreciable plastic deformation ...
  7. [7]
    [PDF] Introduction to Fracture Mechanics - MIT
    Jun 14, 2001 · The term “fracture mechanics” refers to a vital specialization within solid mechanics in which the presence of a crack is assumed, and we wish ...
  8. [8]
    VI. The phenomena of rupture and flow in solids - Journals
    According to these hypotheses rupture may be expected if (a) the maximum tensile stress, (b) the maximum extension, exceeds a certain critical value. Moreover, ...
  9. [9]
    [PDF] The Phenomena of Rupture and Flow in Solids - AA Griffith
    of the two hypotheses of rupture commonly used for solids which are elastic to fracture. ... first sight inexplicable from the standpoint of the molecular theory, ...
  10. [10]
    The work of fracture and its measurement in metals, ceramics and ...
    A method of measuring the work of fracture is described and assessed. Typical values for a number of materials are given and various mechanisms for the ene.
  11. [11]
    ADM guidance—Ceramics: Fracture toughness testing and method ...
    Ceramics and glasses are included in the first category and hence are regarded as brittle solids, whereas most metals and polymers belong to the latter category ...
  12. [12]
    Analysis of Stresses and Strains Near the End of a Crack Traversing ...
    Jun 10, 2021 · Analysis of Stresses and Strains Near the End of a Crack Traversing a Plate Available. G. R. Irwin.
  13. [13]
    [PDF] A Path Independent Integral and the Approximate Analysis of Strain ...
    A path-independent integral is a line integral that has the same value for all paths around a notch tip, used to analyze strain concentration.
  14. [14]
    Mechanical Properties of Materials | MechaniCalc
    The ultimate strength is also referred to as the tensile strength. After ... Additionally, the ultimate strength is coincident with the fracture point.
  15. [15]
    Understanding the 3 Types of Tensile Strength - Corrosionpedia
    May 11, 2020 · The yield, ultimate and fracture strength of materials are essential engineering properties that help determine how components will perform when subjected to ...
  16. [16]
    [PDF] THEORETICAL STRENGTH OF MATERIALS - DTIC
    Consequently, considerations of theoretical strength are particularly pertinent for them. The relations between theoretical strength estimates and the behavior.
  17. [17]
    [PDF] Temperature and strain-rate dependent fracture strength of graphene
    Sep 22, 2010 · The fracture strength ␴r is defined as the stress in the structure when breakdown occurs. With constant strain rate, ␴共t兲 is directly related ...
  18. [18]
    [PDF] A practical and systematic review of Weibull statistics for reporting ...
    By 1977, Jayatilaka and Trustrum [16] used fracture mechanics to develop a general expression for the failure probability using several general flaw size ...
  19. [19]
    Corrosion Effects on Fracture Toughness Properties of Wire Arc ...
    According to this study, the ultimate load and fracture toughness decreased for the corroded specimens. Additionally, it was established that the increase in ...
  20. [20]
    Understanding and mitigating hydrogen embrittlement of steels - NIH
    It can significantly reduce the ductility and load-bearing capacity and cause cracking and catastrophic brittle failures at stresses below the yield stress of ...
  21. [21]
    [PDF] Griffith theory of brittle fracture
    The Griffith equation is strongly dependent on the crack size a,and ... • The criterion for a material to change its fracture behaviour from ductile.
  22. [22]
    E399 Standard Test Method for Linear-Elastic Plane-Strain Fracture ...
    Aug 16, 2019 · 5.1 The property KIc determined by this test method characterizes the resistance of a material to fracture in a neutral environment in the ...
  23. [23]
    J-integral and crack driving force in elastic–plastic materials
    This paper discusses the crack driving force in elastic–plastic materials, with particular emphasis on incremental plasticity.Missing: original | Show results with:original
  24. [24]
    Ductile Fracture - an overview | ScienceDirect Topics
    Ductile fracture is defined as the result of a damage process by plasticity-controlled void nucleation, growth, and coalescence.
  25. [25]
    On localization and void coalescence as a precursor to ductile fracture
    Mar 28, 2015 · Two modes of plastic flow localization commonly occur in the ductile fracture of structural metals undergoing damage and failure by the ...
  26. [26]
    [PDF] On localization and void coalescence as a precursor to ductile fracture
    Feb 25, 2015 · Two modes of plastic flow localization commonly occur in the ductile fracture of structural metals undergoing damage and failure by the ...
  27. [27]
    Ductile-to-Brittle Transition Temperature - ScienceDirect.com
    Ductile-to-brittle transition temperature (DBTT) is the temperature below which materials exhibit low ductility and higher brittleness, while above this  ...
  28. [28]
    Ductile-brittle transition temperature - DoITPoMS
    The ductile-brittle transition temperature can be found by examining the material for a range of temperatures using the Charpy impact test.
  29. [29]
    [PDF] Scanning Electron Microscope
    Brittle fractures usually propagate by either or both of two fracture modes: cleavage and intergranular. In most cases it is necessary to study the fracture ...
  30. [30]
    [PDF] The Transferability of Brittle Fracture Toughness Characteristics
    CLEAVAGE FRACTURE AND WEIBULL STRESS. The transgranular cleavage fracture is a sequential process involving crack initiation and propagation. In most steels.
  31. [31]
    Relation of ductile-to-brittle transition temperature to phosphorus ...
    ▻ A relationship between DBTT and phosphorus segregation is established. ▻ The DBTT increases linearly with increasing phosphorus boundary concentration. ▻ Cold ...
  32. [32]
    [PDF] Lack of wonder and openness lead to failures or at least failed to ...
    Study was conducted to determine failure cause (1943-1946). •~5000 Merchant ... 6) Most failures were on Liberty ships. 7) Every fracture originated at a ...
  33. [33]
    A reconciliation of dynamic crack velocity and Rayleigh wave speed ...
    A reconciliation of dynamic crack velocity and Rayleigh wave speed in isotropic brittle solids ... fracture speeds comparable to the Rayleigh wave speed.
  34. [34]
    [PDF] Fractography of Ceramics and Glasses, Third Edition
    Fractography is a powerful but underutilized tool for the analysis of fractured glasses and ceramics. It may be applied to fractures in the laboratory or for.
  35. [35]
    New insights from crystallography into the effect of Ni content on ...
    The results indicated that increasing the Ni content can reduce the DBTT of HSLA steel and improve the impact toughness at low temperatures.
  36. [36]
    Ductile Fracture - an overview | ScienceDirect Topics
    We will reserve the term 'ductile fracture' for the process of damage nucleation followed by a phase of damage growth and coalescence driven by plastic ...
  37. [37]
    Void-Induced Ductile Fracture of Metals: Experimental Observations
    The paper presents a literature review on the development of microvoids in metals, leading to ductile fracture associated with plastic deformation.
  38. [38]
    Ductile Fracture - an overview | ScienceDirect Topics
    Consequently, the tangling of dislocations results in significant work hardening, leading to a very rapid increase in the flow stress and further increases in ...
  39. [39]
    Four basic types of fracture mechanisms - Gear Solutions Magazine
    Jan 15, 2020 · Figure 2: Schematic representation of the creation of micro-void coalescence (dimples) of a loaded member. Ductile failure can occur with any of ...
  40. [40]
    [PDF] Fracture of Fillet Welds Under Extreme Loading - DSpace@MIT
    Fracture was observed to take place either in the longitudinal fillet tearing mode or in the opening mode at the toe of the fillet weld. The tearing resistance ...
  41. [41]
    [PDF] chapter 3: macro/micro - aspects of fatigue of metals - EFatigue
    ▫ Beach marks indicative of crack growth. ▫ Distinct final fracture region. ▫ Representative macroscopic fatigue fracture surfaces. Page ...Missing: morphology | Show results with:morphology
  42. [42]
    [PDF] Failure Analysis and Fracture
    Marks. A fracture surface that is smoother and possibly discolored. Ratchet marks when multiple origins are present. Shear lip. Page 42. 42. Fatigue Cracking. A ...
  43. [43]
    Ductile Failure - an overview | ScienceDirect Topics
    This type of failure typically exhibits rough and torn fracture surfaces, with shear surfaces angled at 45° to the applied load. AI generated definition based ...
  44. [44]
    [PDF] Measurement Good Practice Guide No. 15 Fractography of Brittle ...
    mist/hackle boundaries are very clear, which would permit a good fracture mechanical assessment of the fracture process. Using the SEM, the origin can be ...
  45. [45]
    The Forces Exerted on Dislocations and the Stress Fields Produced ...
    This research constitutes a portion of a thesis submitted by M. O. Peach in partial fulfillment of the requirements for the degree of Doctor of Science at ...
  46. [46]
    [PDF] Ductile versus brittle behaviour of crystalst
    A necessary criterion for brittle fro.oture in crystals is established in terms of tho spontaneous emission of dislocations from nn o.tomioally sho.rp cleavage.
  47. [47]
    [PDF] INTERGRANULAR FRACTURE IN METALS - HAL
    Feb 4, 2008 · This paper concerns brittle intergranular fracture in metals. It is shown that the primary cause of intergranular fracture is impurity ...
  48. [48]
    Nanoscale ductile fracture and associated atomistic mechanisms in ...
    Sep 8, 2023 · These dislocation activities give rise to an alternating sequence of crack-tip plastic shearing, resulting in crack blunting, and local ...Results · In Situ Hrtem Imaging · Md Simulation
  49. [49]
    Deformation twinning in body-centered cubic metals and alloys
    In this review, we systematically summarize recent advances of deformation twinning in BCC metals and alloys in past few decades.Missing: transformations seminal
  50. [50]
    The dual role of martensitic transformation in fatigue crack growth
    Feb 24, 2022 · Here, we identify two antagonistic mechanisms mediated by martensitic transformation during the fatigue process through in situ observations.
  51. [51]
    Standard Test Method for Measurement of Fracture Toughness - ASTM
    Oct 24, 2022 · Toughness can be measured in the R-curve format or as a point value. The fracture toughness determined in accordance with this test method is ...
  52. [52]
    Evaluation of dynamic fracture toughness K Id by Hopkinson ...
    The Hopkinson pressure bar experimental technique has been proposed as an improvement on the instrumented Charpy impact test at loading rates over 106 MPa m / s ...
  53. [53]
    Size effect and other effects on mode I fracture toughness using two ...
    Mode I fracture toughness (KIC) is affected by specimen size, with a greater size effect in SCB tests. pCT tests show more consistent KIC values, and the ...
  54. [54]
    Statistical and Constraint Loss Size Effects on Cleavage Fracture ...
    A systematic investigation of the effects of specimen size on the cleavage fracture toughness of a typical pressure vessel steel is reported.
  55. [55]
    [PDF] 7. fracture
    Fracture can be defined as the process of separation (or fragmentation) of a solid into two or more parts under the action of a stress. So-defined, fracture ...<|separator|>
  56. [56]
    [PDF] Surface Flaw Reliability Analysis of Ceramic Components With the ...
    Statistical fast fracture models, based on weakest link theories (WLT) of Weibull, have been previously developed for both. tyDes of flaw populations (Weibull, ...
  57. [57]
    characteristic strength and fracture toughness - ScienceDirect.com
    The Weibull modulus M indicates the degree of the fracture strength σf scatters, which is typically between 5 and 15 for polycrystalline ceramics. A smaller M ...
  58. [58]
    Thermal Shock - an overview | ScienceDirect Topics
    ... thermal shock resistance parameter, R. (8) R = σ 1 − ν α E. The R parameter estimates the Δ T (in °C) that the material can withstand prior to formation of ...
  59. [59]
    Thermal Shock Resistance - Morgan Technical Ceramics
    Thermal shock resistance is a material's ability to withstand rapid temperature changes. Rapid cooling causes tensile stress, potentially cracking the ceramic.
  60. [60]
    Ceramics for Dental Applications: A Review - PMC - NIH
    Jan 11, 2010 · Unfortunately, pores are not the only defects found in veneering ceramics, as shown in Figure 2, cracks and inclusions are also present. The ...
  61. [61]
    Visualization of transformation toughening of zirconia ceramics ...
    Jun 7, 2021 · Transformation toughening, where phase transformations around the crack tip reduce the crack-driving force, is one of the main extrinsic ...
  62. [62]
    [PDF] Science of zirconia-related engineering ceramics
    Since the publication of the seminal paper on toughened zirconia by Garvie et al. (1975), this material has become the most important ceramic material of the ...
  63. [63]
    [PDF] MECHANISMS OF TOUGHENING IN CERAMICS - Harvard University
    Toughening mechanisms in ceramics include transformation toughening, toughening by metallic particles, and fiber reinforcement. Transformation toughening uses ...
  64. [64]
    Stress Corrosion and Static Fatigue of Glass - WIEDERHORN - 1970
    Stress corrosion cracking of six glasses was studied using fracture mechanics techniques. Crack velocities in water were measured as a function of applied ...Missing: tiles windows
  65. [65]
    A review on micromechanical modelling of progressive failure in ...
    Five microscopic main failure modes, namely fibre breakage, fibre pull-out, fibre/matrix debonding, matrix cracking and delamination are all captured in Fig. 1, ...
  66. [66]
    Mode I fracture toughness of fiber-reinforced polymer composites
    Jun 24, 2019 · This review paper emphasizes on the effects of different reinforcement structures on mode I fracture toughness and possible ways to improve fracture toughness.
  67. [67]
    Mode I Fatigue of Fibre Reinforced Polymeric Composites: A Review
    Fibre fracture, matrix fracture and fibre–matrix delamination are the three dominant fatigue failure modes observed in unidirectional single layer composite.Delamination Modes · 2.3. 3d Woven Composites · 3. Failure Modes In...
  68. [68]
    Fibre Bundle Model - an overview | ScienceDirect Topics
    Fibre bundle models (FBM) are one of the most important theoretical approaches to the damage and fracture of fibre-reinforced composite materials which have ...
  69. [69]
    Failure processes in elastic fiber bundles | Rev. Mod. Phys.
    Mar 1, 2010 · The fiber bundle model describes a collection of elastic fibers under load. The fibers fail successively and, for each failure, the load distribution among the ...
  70. [70]
    Modeling of fiber toughening in fiber-reinforced ceramic composites
    When a macroscopic crack propagates in ceramic matrix, some fibers with higher strength will bridge the crack, and some fibers with lower strength will break ...
  71. [71]
    [PDF] Micromechanics Analysis of Fiber Toughening in Ceramic-Matrix ...
    It is found that crack bridging is the dominant mechanism and the toughening in the bridged composite is strongly dependent on stiffness ratio, crack tip ...
  72. [72]
    Toughening of ceramics and ceramic composites through ...
    Jan 2, 2025 · Fracture mechanics has always played a major role in our understanding of the structural properties of ceramics and how to design them for ...
  73. [73]
    Rule-of-Mixture Equation - an overview | ScienceDirect Topics
    The rule of mixtures equation refers to a set of equations used to predict the properties, such as modulus and strength, of fiber-reinforced composites ...
  74. [74]
    [PDF] Fracture Characteristics of Fiber Composites, - DTIC
    Perhaps the first model to express composite strength as a function of the reinforcing fibers was the rule of mixtures, ROM. This model attempted to predict a ...
  75. [75]
    [PDF] Micromechanical failure in fiber-reinforced composites
    Mar 31, 2014 · that with weak interfaces the fiber/matrix debonding mainly controlls failure of composites while the matrix deformation is the dominant ...
  76. [76]
    [PDF] fracture toughness of carbon fiber composites containing various ...
    For example, a self- healing composite that possesses aerospace quality consolidation with fiber volume fraction. (FVF)≈60% and void volume fraction (VVF) < 2% ...
  77. [77]
    Bundle pullout—a failure mechanism limiting the tensile strength of ...
    A phenomenon called “bundle pullout”; a concentration of fiber failures in a small volume element enables whole clusters of fibers and matrix to be pulled out.Missing: tow | Show results with:tow
  78. [78]
    Brittle Fracture: When Ships Split in Two - Mariners' Museum
    Jan 28, 2021 · Brittle fracture, caused by impurities like sulfur and phosphorus in steel, makes ships prone to splitting due to welding issues and stress ...
  79. [79]
    De Havilland DH-106 Comet 1 | Federal Aviation Administration
    Mar 7, 2023 · The fuselage finally failed at 16,000 cycles due to fatigue cracks at the corner of a squarish cabin window. ... First, after a summary of the ...
  80. [80]
    The Hyatt Regency Walkway Collapse - ASCE
    Jan 1, 2007 · The collapse was traced to failure of the connections between the fourth-story box beams and the hanger rods supporting the second-story and ...
  81. [81]
    Pacific Gas & Electric Pipeline Rupture in San Bruno, CA | PHMSA
    On September 9, 2010 at 6:11 pm, a 30 inch diameter natural gas transmission pipeline in San Bruno, CA ruptured and released vast quantities of natural gas.
  82. [82]
    Five Years After Deadly San Bruno Explosion: Are We Safer? - KQED
    Sep 8, 2015 · With Line 132 carrying gas at high pressure, the crack grew year after year until it was ready to fail. The pipeline was never pressure tested.
  83. [83]
    [PDF] EVOLUTION OF ENGINEERING PRINCIPLES FOR FRACTURE ...
    Fracture-safe design implies deliberate engineering analyses of the possibilities for structural failure by fracture-in addition to the usual considerations ...
  84. [84]
    [PDF] Nondestructive Testing and Fracture Mechanics - NDT.net
    Jun 1, 2016 · NDT and fracture mechanics are used for component integrity evaluation. NDT methods provide defect size data, which is essential for fracture ...