Fact-checked by Grok 2 weeks ago

Fracture mechanics

Fracture mechanics is a specialized branch of that analyzes the behavior of in materials, establishing quantitative relationships between length, the material's resistance to propagation, and the applied stresses that lead to rapid and structural failure. It assumes the presence of preexisting flaws or and focuses on predicting their growth under load, which is crucial for designing safe structures in materials where small defects can cause even at relatively low nominal stresses. The field originated with A. A. Griffith's seminal 1921 work on brittle fracture in glass, where he introduced an energy-balance criterion showing that crack propagation occurs when the decrease in equals or exceeds the energy required to create new crack surfaces. In the 1940s and 1950s, George R. Irwin extended this theory to ductile materials by incorporating plastic energy dissipation and developing the concept of the (K), a parameter that characterizes the stress state near the crack tip and scales with applied load and crack size. Irwin's modifications led to linear elastic fracture mechanics (LEFM), which assumes small-scale yielding at the crack tip and uses elastic theory to predict fracture. Central to fracture mechanics are material properties like fracture toughness (K_Ic for mode I tensile loading), defined as the critical stress intensity factor beyond which a advances unstably, and the strain energy release rate (G), which quantifies the energy available for growth per unit area. For instance, in plane strain conditions, K_Ic relates to G_c via K_Ic = √(E G_c / (1 - ν²)), where E is the and ν is , enabling comparisons across materials like metals (K_Ic ≈ 50–150 MPa√m) and ceramics (lower values). These parameters allow engineers to assess critical flaw sizes and ensure components withstand service loads without . In engineering practice, fracture mechanics underpins damage tolerance design, particularly in , where it predicts crack growth, in composites, and residual strength under monotonic or cyclic loading to certify airframes for safe operation. Applications extend to civil structures, pressure vessels, and biomedical implants, emphasizing the need for nondestructive and to mitigate flaw-induced risks, with annual economic impacts from fractures estimated in billions.

Introduction and Fundamentals

Historical Development

Early observations of brittle fracture in materials such as and metals date back to the late 19th and early 20th centuries, with researchers noting unexpected weaknesses due to flaws or cracks that concentrated es far beyond nominal levels. A pivotal early contribution came from C. E. Inglis in , who analyzed concentrations around elliptical holes in thin plates under , demonstrating that the maximum at the hole's edge increases sharply with the ratio of the ellipse's semi-major axis to its , providing the first theoretical basis for crack-like flaws as stress raisers. In 1920–1921, A. A. Griffith built on Inglis's work through experiments on glass fibers and plates, observing that surface scratches reduced tensile strength by factors of up to four due to amplified stresses at flaw tips. Griffith introduced the foundational concept of an energy balance for crack propagation in brittle materials, positing that a crack advances only when the decrease in exceeds the energy required to create new surfaces, as detailed in his seminal papers published in the Philosophical Transactions of the Royal Society. During , investigations into aircraft and armor plate failures highlighted the limitations of Griffith's brittle fracture theory for ductile metals, prompting George R. Irwin at the U.S. Naval Research Laboratory to extend it in the 1940s and 1950s. Irwin's work, motivated by brittle fractures in high-strength steels used in military applications, shifted the focus from global energy release to local stress fields near the crack tip, introducing the as a key parameter characterizing crack-tip stress states. Fracture mechanics gained rapid adoption in the 1950s, particularly in following incidents like the 1954 jet disasters, which underscored the need for crack-tolerant design in aircraft structures. By the , standardization efforts advanced with the development of testing protocols, culminating in the American Society for Testing and Materials (ASTM) standard E399 for plane-strain (K_Ic) measurement, first issued as a tentative method in but rooted in research. A major milestone was the First International Conference on Fracture in , , in 1965, organized by T. Yokobori, T. Kawasaki, and J. L. Swedlow, which fostered global collaboration and marked the field's emergence as a distinct discipline.

Basic Concepts and Prerequisites

Fracture mechanics analyzes the behavior of in materials under load, focusing on how cracks initiate, propagate, and lead to . Central to this field are the definitions of crack loading modes, which describe the relative displacement of crack surfaces. There are three primary modes: Mode I, the opening or tensile mode, where the crack surfaces separate to the crack plane; Mode II, the sliding or in-plane mode, where the crack surfaces slide over each other parallel to the crack plane and the of crack propagation; and Mode III, the tearing or out-of-plane mode, where the crack surfaces slide relative to each other in a to the crack plane and propagation . These modes capture the essential of crack deformation and are foundational for modeling crack-tip conditions. Fracture toughness quantifies a 's resistance to growth, representing the critical condition at which a becomes unstable and propagates rapidly. It is typically measured under controlled conditions, such as plane strain in Mode I loading, and serves as a key for assessing structural . Near the crack tip, stress and strain fields exhibit a characteristic singularity, where stresses increase dramatically as the distance from the tip approaches zero. In polar coordinates (r, \theta) centered at the crack tip, with r as the radial distance and \theta as the angular position, the stress components follow an asymptotic form dominated by the inverse square root dependence on r. Specifically, the stresses scale as \sigma \sim \frac{1}{\sqrt{r}}, leading to infinite values at r = 0 in the ideal linear elastic model. This singularity highlights the intense local deformation at the crack tip, which drives propagation, though real materials mitigate it through atomic-scale effects or plasticity. Strains exhibit a similar \sqrt{r} singularity in displacement fields. Fracture mechanics relies on several material assumptions to simplify analysis while capturing essential behavior. Materials are typically assumed isotropic, meaning properties are direction-independent, and homogeneous, with uniform composition throughout. Small-scale yielding is presupposed, where plastic deformation is confined to a small region near the crack tip, allowing linear elastic theory to approximate the surrounding field. Two loading conditions are distinguished: plane stress, applicable to thin specimens where out-of-plane stresses are negligible, and plane strain, relevant for thick components where constrained deformation leads to triaxiality at the crack tip. Geometric factors play a crucial role in crack behavior, particularly the crack length a, which influences stress amplification. In an infinite plate, the effective crack size is often taken as the half-length $2a for a central crack or the full length for edge cracks, determining the remote stress required for propagation. Defects, such as inclusions or manufacturing flaws, act as stress concentrators, creating local peaks that can initiate cracks by magnifying applied stresses by factors depending on defect geometry. Prerequisite concepts from underpin fracture analysis. governs the linear elastic response, stating that \sigma is proportional to \epsilon via the E, as \sigma = E \epsilon, ensuring reversible deformation away from the crack. justifies treating far-field stresses as uniform, asserting that the exact load distribution details become irrelevant at distances much larger than the loaded region's size, allowing focus on the crack-tip singularity without detailed boundary effects. These basics enable the energy balance ideas, as later applied by Griffith, to predict brittle fracture thresholds.

Linear Elastic Fracture Mechanics

Griffith's Criterion

Griffith's criterion provides the foundational energy-based theory for predicting brittle in elastic materials, establishing the condition under which a will propagate spontaneously. Developed by A. A. Griffith in 1921, the criterion posits that occurs when the release of stored elastic during extension exactly balances the energy required to create new surfaces. This approach resolves the discrepancy between theoretical strengths of materials based on atomic bond energies and their observed lower strengths due to microscopic flaws acting as stress concentrators. The core of the theory lies in the energy balance for crack growth. Consider a cracked body under load; the total potential energy U consists of the elastic strain energy stored in the material and the surface energy associated with the crack faces. As the crack advances, elastic energy is released, but surface energy is consumed to form new free surfaces. For a unit thickness, the change in total energy is given by \Delta U = U_{\text{elastic}} + U_{\text{surface}}, where U_{\text{surface}} = 2 \gamma \cdot 2a for a central crack of length $2a (accounting for two crack faces, with \gamma as the surface energy per unit area). The elastic strain energy released is negative, reflecting the reduction in stored energy. Equilibrium is achieved when the derivative \frac{dU}{da} = 0, marking the onset of crack propagation; beyond this point, if \frac{dU}{da} < 0, the crack grows unstably as the energy release exceeds the surface energy requirement. For the specific case of an infinite plate under uniform tensile stress \sigma containing a central through-crack of length $2a, Griffith derived the explicit form of this balance assuming plane stress conditions. The released elastic strain energy is U_{\text{elastic}} = -\frac{\pi \sigma^2 a^2}{E}, where E is , derived from the stress perturbation around the crack using Inglis' elliptical hole solution adapted for a sharp crack. The surface energy term is U_{\text{surface}} = 4 \gamma a. Setting \frac{dU}{da} = 0 yields the critical stress for fracture: \sigma_c = \sqrt{\frac{2 E \gamma}{\pi a}} This formula indicates that the fracture stress decreases with increasing crack length a and increases with material stiffness E and surface energy \gamma, highlighting the role of flaws in reducing strength. For unstable growth, the second derivative \frac{d^2 U}{da^2} = -\frac{2 \pi \sigma^2}{E} < 0 confirms the equilibrium is a maximum, leading to rapid propagation once initiated. The criterion applies primarily to perfectly brittle materials, such as glass, where crack propagation is catastrophic without significant plastic deformation. In glass, microscopic surface scratches or internal flaws serve as preexisting cracks, and the theory explains why polished or flaw-free specimens exhibit strengths approaching theoretical values, while flawed ones fail at much lower stresses. Griffith demonstrated this through experiments on glass fibers approximately 2 inches long and diameters ranging from 0.002 to 0.01 inches, measuring tensile strengths that aligned closely with the predicted \sigma_c. For instance, using E \approx 9.01 \times 10^6 psi and inferred \gamma \approx 350 erg/cm² from rupture data, the observed fracture stresses matched the formula within experimental error, validating the energy balance for flaw-induced failure in brittle solids. These tests showed that finer fibers, with presumably smaller flaws, sustained higher stresses, consistent with the inverse square-root dependence on crack size. Key assumptions underlying the criterion include linear elastic behavior throughout the material, negligible plasticity or energy dissipation at the crack tip, and infinitesimal crack sizes relative to the body dimensions to justify the infinite plate approximation. The theory presumes ideally brittle fracture without atomic-scale details, focusing on continuum mechanics, and applies to mode I (tensile) opening of straight cracks in homogeneous, isotropic media.

Irwin's Modifications and Stress Intensity Factor

George R. Irwin advanced Griffith's energy-based theory of brittle fracture by developing a local stress analysis near the crack tip, providing a more versatile parameter for predicting fracture in engineering applications. Motivated by the need to characterize the singular stress field dominating crack propagation, Irwin focused on the asymptotic behavior close to the crack tip, where stresses intensify inversely with the square root of the distance from the tip. This approach allowed for a scalable measure of crack severity that could be applied across different geometries and loading conditions, bridging theoretical insights with practical design needs. In his foundational 1957 analysis of stresses and strains near the end of a crack traversing an infinite plate, Irwin derived the characteristic near-tip stress field for mode I (tensile opening) loading. The normal stress component perpendicular to the crack plane, σ_yy, is expressed as \sigma_{yy} = \frac{K_I}{\sqrt{2\pi r}} \cos\left(\frac{\theta}{2}\right) \left[1 + \sin\left(\frac{\theta}{2}\right) \sin(\theta)\right], where r is the radial distance from the crack tip, θ is the polar angle measured from the crack plane ahead, and K_I is the mode I stress intensity factor that quantifies the amplitude of the singularity. This formulation reveals that the stress field near the tip is self-similar and determined solely by K_I, independent of remote boundary conditions beyond a certain distance. Irwin's derivation confirmed the 1/√r singularity predicted by elasticity theory for cracks, enabling engineers to assess local conditions without solving the full boundary value problem. The stress intensity factor K_I is defined for a central through-crack of length 2a in an infinite plate subjected to uniform remote tensile stress σ as K_I = σ √(πa). For finite geometries, incorporated a dimensionless geometry factor Y to generalize the expression: K_I = Y σ √(πa), where Y = 1 for the infinite plate but adjusts for boundary influences, such as finite width, to maintain accuracy in stress field predictions. This relates briefly to σ_c for fracture via K = σ √(πa), linking the local stress parameter to the global energy criterion. Irwin extended the stress intensity factor concept to mixed-mode fracture, introducing K_II for in-plane shear (sliding mode) and K_III for anti-plane shear (tearing mode), each governing the corresponding singular stress components near the tip. These factors together fully describe the three-dimensional stress state at the crack front under combined loadings. The material's resistance to fracture is characterized by the critical stress intensity factor K_c, a key property measuring the value of K at unstable crack growth initiation, independent of specimen size under valid linear elastic conditions. In the same 1957 work, Irwin outlined modifications to the stress intensity factor to address finite width effects and three-dimensional influences, such as thickness variations or non-through cracks, by refining Y to incorporate these real-world deviations from the ideal infinite plate assumption. These adjustments ensured the parameter's applicability to practical components like plates and beams with bounded dimensions. A primary benefit of Irwin's stress intensity factor is its inherent scalability, as K varies linearly with applied stress and with the square root of crack length, permitting the superposition of multiple loads or geometric features to compute the total intensity at the tip. This property simplifies fracture assessments in complex structures, such as aircraft components, by allowing modular analysis of crack configurations.

Strain Energy Release Rate

The strain energy release rate, denoted G, quantifies the driving force for crack propagation in linear elastic fracture mechanics by measuring the energy made available for creating new crack surfaces per unit increase in crack area. It is defined as G = -\frac{\partial \Pi}{\partial A}, where \Pi is the total potential energy of the loaded system and A is the crack surface area; for a through-crack of length a in a specimen of unit thickness, this reduces to G = -\frac{d\Pi}{da}. In practice, under constant load P, G corresponds to the rate of strain energy release, G = \frac{dU}{da}, where U is the elastic strain energy stored in the body. This energy-based criterion extends the original balance proposed by to arbitrary crack geometries and loading configurations, providing a global measure independent of local stress fields near the crack tip. A practical method for computing G involves the compliance of the structure, defined as C = \frac{\delta}{P}, where \delta is the load-point displacement. The stored strain energy is then U = \frac{1}{2} P^2 C, leading to G = \frac{P^2}{2B} \frac{dC}{da} for a specimen of width B, applicable under both constant load and constant displacement boundary conditions. This compliance technique, introduced by in the mid-1950s, facilitates experimental determination of G by measuring load-displacement curves for incrementally extended cracks. For brittle materials, unstable crack growth initiates when G reaches the critical value G_c = 2\gamma, where \gamma is the specific surface energy required to create the fracture surfaces. The strain energy release rate G is equivalently expressed in terms of the mode I stress intensity factor K_I as G_I = \frac{K_I^2}{E} under plane stress conditions and G_I = \frac{K_I^2 (1 - \nu^2)}{E} under plane strain conditions, where E is the and \nu is ; similar relations hold for modes II and III, with total G = G_I + G_{II} + G_{III}. Here, K_I provides a complementary stress-based characterization of the crack tip loading that scales with geometry and correlates directly with G. In conservative elastic fields, G exhibits path independence, meaning its value remains unchanged regardless of the integration contour surrounding the crack tip, as long as the path encloses only the crack singularity—this property enables robust numerical evaluations. The SERR concept finds wide application in assessing collinear crack extension, where multiple aligned cracks propagate sequentially, and embedded flaws, such as internal voids or inclusions that evolve into cracks under load. In complex structures, finite element analysis is routinely used to compute G via virtual crack advance techniques, which estimate energy changes from nodal displacements, or through equivalent domain integrals that approximate the path-independent contour. Developments in the 1960s by G. C. Sih and H. Liebowitz extended these methods to multi-mode scenarios, formulating expressions for G under combined opening, sliding, and tearing modes to predict fracture in anisotropic and composite materials.

Crack Tip Plastic Zone Adjustment

In linear elastic fracture mechanics, the idealization of purely elastic material behavior fails near the crack tip, where stresses exceed the yield strength, leading to localized plastic deformation in ductile materials. The crack tip plastic zone adjustment modifies parameters to account for this small-scale yielding, enabling more accurate predictions of fracture behavior without departing from the elastic framework. These corrections recognize that plasticity relaxes stresses at the tip, effectively extending the crack length and altering the stress field ahead of the zone. The size of the plastic zone, denoted as r_p, is estimated in Irwin's model for plane stress conditions by the approximate relation r_p \approx \frac{1}{2\pi} \left( \frac{K_I}{\sigma_y} \right)^2, where K_I is the mode I stress intensity factor and \sigma_y is the material's yield strength. This formula derives from the point where the elastic stress singularity reaches \sigma_y, assuming elastic-perfectly plastic behavior. The shape of the zone is influenced by the von Mises yield criterion, which yields a roughly circular region in plane stress for isotropic metals. To incorporate this plasticity into LEFM calculations, Irwin introduced an effective crack length correction: a_{\text{eff}} = a + r_p / 2, where a is the observed physical crack length. This adjustment positions the singularity at the boundary of the plastic zone, better representing the elastic stress distribution beyond it. The corresponding effective stress intensity factor then becomes K_{\text{eff}} = \sigma \sqrt{\pi a_{\text{eff}}}, with \sigma as the remote applied stress, and similarly adjusts the strain energy release rate G. These modifications enhance the accuracy of fracture toughness assessments in mildly ductile alloys. A complementary model, the Dugdale strip yield approach, treats the plastic zone as a narrow strip ahead of the crack where material yields at constant \sigma_y, creating a cohesive zone that shields the tip singularity. This bar-model solution for infinite plates predicts a plastic zone size comparable to Irwin's estimate and is especially applicable to thin metallic sheets under tension. Such adjustments apply under small-scale yielding conditions, requiring r_p to be much smaller than the specimen width, crack length, and remaining ligament (typically r_p less than 1/25 of these dimensions) to confine plasticity and preserve LEFM validity. Early experimental validation came from optical microscopy studies in the 1950s and 1960s on metals like aluminum alloys and steels, which visualized slip lines and etch pits revealing plastic zones matching the predicted sizes and shapes near fatigue cracks.

Limitations and Validity Conditions

Linear elastic fracture mechanics (LEFM) relies on several key assumptions that limit its applicability, primarily the requirement for predominantly elastic material behavior with only localized plasticity at the crack tip. It neglects significant plastic deformation, time-dependent phenomena such as creep or viscoelasticity, and nonlinear material responses, making it unsuitable for materials or conditions where these effects dominate. LEFM is valid only when the stress intensity factor K remains below the critical fracture toughness K_c, ensuring crack stability, and when the plastic zone size r_p is small relative to the crack length a, typically r_p / a < 0.05. Validity criteria for LEFM include monotonic loading to avoid cyclic effects like fatigue, quasi-static crack propagation to exclude dynamic influences, and homogeneous, isotropic materials to maintain the assumed stress field singularity. Irwin's plastic zone criterion further specifies that the plastic zone must be contained within the specimen, with r_p much smaller than the ligament size b (the uncracked region behind the crack), often requiring r_p / b < 0.05 to ensure small-scale yielding conditions hold across the structure. These conditions are typically met in brittle materials under low-temperature or high-strain-rate scenarios, but deviations invalidate the linear elastic framework. Common pitfalls arise in applying LEFM to inappropriate materials, such as overestimation of fracture toughness in ductile metals when small specimens yield plane-stress conditions, leading to apparent toughness values higher than the intrinsic plane-strain K_{Ic} and potentially unconservative designs. Conversely, in composites, LEFM often underestimates toughness due to unaccounted nonlinear effects like fiber bridging or matrix cracking, which enhance resistance beyond elastic predictions. Transition indicators from LEFM validity occur when applied stresses exceed the yield strength \sigma > \sigma_y, promoting extensive , or when release rate G surpasses values dominated by plastic dissipation rather than recovery. Historical cases, such as the failures, illustrate these limits; over 400 ships suffered brittle due to welds and low-temperature loss, where traditional stress analysis failed, but LEFM's assumptions proved inadequate for the variable in hull steels, necessitating broader fracture approaches.

Elastic-Plastic Fracture Mechanics

Crack Tip Opening Displacement

The (CTOD), denoted as δ, quantifies the deformation at the crack tip in materials exhibiting significant plastic behavior, serving as a key parameter in elastic-plastic fracture mechanics for assessing in ductile materials. It is defined as the distance between the crack faces measured normal to the crack plane and extrapolated to the original crack tip position. In test specimens, such as the single-edge notched bend (SENB), CTOD relates to the that forms due to rotation in the ligament ahead of the crack, capturing the extent of crack tip blunting and before unstable . In the 1960s, A.A. Wells formulated CTOD as a fracture criterion, proposing an expression for small-scale yielding conditions where the total CTOD, δ_t, approximates the elastic contribution as δ_t ≈ K_I^2 / (E σ_y) for plane stress, with K_I the mode I stress intensity factor, E the Young's modulus, and σ_y the yield strength; this relation links the linear elastic parameter K to plastic deformation at the tip. The plastic component, δ_p, derives from the Dugdale bar model of cohesive yielding along a plastic zone ahead of the crack, providing a more accurate representation for contained yielding where δ_p contributes significantly to the total displacement. The critical CTOD, δ_c, at the onset of crack extension then serves as a material toughness parameter analogous to the critical stress intensity factor K_{Ic} but valid in the presence of plasticity. CTOD is typically measured experimentally using a clip gauge mounted across the crack mouth or at knife-edge slots on fracture toughness specimens like compact tension (CT) or SENB geometries, as standardized in ASTM E1820, which outlines procedures for determining critical CTOD values at crack initiation or extension under displacement-controlled loading. The gauge records the crack opening displacement (COD) versus load curve, from which CTOD is calculated via extrapolation or plastic hinge rotation factors to estimate the opening at the reference point. This method ensures reproducibility and accounts for both elastic and plastic contributions in testing. Compared to the stress intensity factor K, CTOD offers advantages by directly incorporating crack tip plasticity as a physical measure of deformation, making it suitable for applications in ductile steels where linear assumptions fail; it has been extensively applied in the assessment of for offshore oil and gas structures, such as pipelines and platforms, to predict stable growth and arrest behavior under high-strain conditions. In some cases, CTOD relates to the through δ ≈ J / (m σ_y), where m is a factor (typically 1.15–2), enabling path-independent characterization of driving force. Despite these benefits, CTOD measurements are sensitive to specimen and crack-tip , with values varying due to differences in triaxiality between test specimens and structural components, often requiring for accurate transferability. Additionally, valid CTOD testing typically demands deep (a/W ≥ 0.5, where a is crack depth and W is specimen width) to ensure sufficient and plane strain conditions, limiting its use in shallow- scenarios without modifications.

J-Integral Approach

The , introduced by James R. Rice in , serves as a path-independent integral that quantifies the crack driving force in nonlinear elastic materials, extending fracture mechanics beyond the linear elastic regime. It is defined mathematically as J = \int_{\Gamma} \left( W \, dy - \mathbf{T} \cdot \frac{\partial \mathbf{u}}{\partial x} \, ds \right), where \Gamma is a closed encircling the crack tip in the deformed configuration, W is the strain energy density, \mathbf{T} is the traction on the contour, \mathbf{u} is the , and ds and dy are elements along the contour and in the y-direction, respectively. This formulation captures the total into the crack tip region, making it applicable to materials exhibiting significant . Key properties of the J-integral include its path independence for hyperelastic materials under quasi-static loading, which holds even in the presence of nonlinearity as long as the deformation is conservative, and its equivalence to the strain energy release rate G in the linear elastic limit. The units of J are energy per unit length (e.g., kJ/m in two-dimensional analyses), reflecting its role as an energy-based measure of crack extension potential. In the linear elastic case, this equivalence J = G allows seamless transition from classical Griffith-Irwin theory to elastic-plastic conditions. For estimation in power-law hardening materials, where the stress-strain response follows \sigma = K \epsilon^n with hardening exponent n, the governs the near-tip fields described by the Hutchinson-Rice-Rosengren (HRR) . These fields predict a distribution of stresses and strains that scale with J, providing a basis for characterizing the size of the plastic zone and the intensity of the crack-tip in elastic-plastic fracture mechanics. The HRR solution, derived independently by Hutchinson, , and Rosengren in 1968-1969, assumes small-scale yielding relative to the crack length and specimen dimensions, enabling quantitative predictions of crack behavior in metals with moderate strain hardening. The critical value J_{Ic}, representing the fracture initiation toughness under plane-strain conditions, is determined experimentally using the ASTM E1820 standard test method, which involves single-specimen unloading techniques on compact or single-edge notched bend specimens to construct J-resistance curves. This approach ensures valid J_{Ic} measurements by verifying proportionality between load-line and extension up to an offset of 0.15 mm. A related parameter is the tip opening (CTOD) \delta, linked to J through the approximate relation J = m \sigma_y \delta, where \sigma_y is the yield stress and m is a constraint factor typically ranging from 1 to 2 depending on geometry and strain hardening. Applications of the proliferated in the 1970s, particularly for assessing fracture in welded structures where heterogeneous material properties and residual stresses complicate linear elastic analyses; it enabled reliable life prediction and fitness-for-service evaluations in pressure vessels and pipelines. adopted the in the mid-1970s for damage-tolerant design of components, integrating it into fracture control plans for space vehicles to account for elastic-plastic behavior under complex loading.

Resistance Curves and Toughness

In fracture mechanics, the resistance curve, or R-curve, quantifies a material's increasing resistance to crack propagation as a function of crack extension, typically plotted as the crack driving force—such as the or —against the Δa. This rising resistance contrasts with the fixed critical values like K_c or J_Ic in linear elastic fracture mechanics, reflecting mechanisms such as crack bridging by uncracked ligaments and plastic stretching of material ahead of the crack tip, which dissipate energy and elevate during stable crack growth. The concept originated with George Irwin's early work on crack extension resistance in the , providing a framework for elastic-plastic regimes where toughness evolves with crack advance. A key parameter derived from the R-curve is the tearing modulus T, which characterizes the stable crack growth rate and is defined as T = \frac{E}{\sigma_y^2} \frac{dJ}{da}, where E is the , σ_y is the yield strength, J is the (used as the driving force on the x-axis), and da is the incremental crack extension. Higher T values indicate greater resistance to unstable tearing, enabling predictions of crack stability in ductile materials under rising load conditions. R-curves are measured experimentally using standardized methods, such as ASTM E561 for metallic materials, which employs center-cracked tension panels or compact tension specimens under monotonic loading to generate K_R curves via techniques or optical crack length measurements. For J_R curves, early multi-specimen techniques involve heat-treating multiple identical specimens to halt crack growth at incremental Δa values, allowing post-test surface to construct the curve, though single-specimen unloading methods later improved efficiency. The R-curve typically begins with an initiation toughness J_Ic at small Δa, rises steeply due to initial plastic zone development, and may plateau at a steady-state value where further mechanisms like void coalescence saturate, influenced by factors such as and microstructure. In ductile alloys, this plateau represents the maximum tearing before unstable fracture. Significant developments in the and focused on J_R curves for ductile alloys like structural steels, with J.A. Begley and J.D. Landes establishing the as a viable resistance parameter in their 1972 ASTM STP 514 paper, enabling quantitative assessment of stable crack growth in engineering applications. Subsequent work by P.C. and colleagues in the late refined tearing instability criteria, applying R-curves to predict ductile in pressure vessels and pipelines, enhancing safety assessments for and components.

Cohesive Zone Modeling

Cohesive zone modeling represents a fundamental approach in fracture mechanics for describing the nonlinear processes at the crack tip, particularly in materials exhibiting some or heterogeneity. Introduced by Barenblatt in the late and early , this framework conceptualizes the fracture process as occurring within a finite "cohesive zone" ahead of the crack tip, where cohesive tractions govern the separation of material surfaces. The core of the model is the traction-separation law, denoted as \sigma(\delta), which relates the cohesive stress \sigma to the crack opening displacement \delta. This law typically features a rising branch to a peak strength \sigma_c, followed by a softening branch until a critical separation \delta_f is reached, beyond which the material fully separates. Barenblatt's formulation aimed to resolve the stress singularity predicted by linear elastic fracture mechanics by distributing the load over this zone, ensuring finite stresses at the crack tip. A notable special case of the cohesive zone model is the Dugdale model, developed in , which simplifies the traction-separation law to a constant yield stress \sigma_y applied over a plastic zone of length R. This zone length is derived as R = \frac{\pi}{8} \left( \frac{K}{\sigma_y} \right)^2, where K is the , providing an analytical estimate for the plastic zone size in ductile materials like metals under small-scale yielding conditions. The Dugdale approach assumes the cohesive zone closes the crack faces to eliminate the , making it particularly useful for bridging linear elastic predictions with plastic effects. In numerical implementations, cohesive zone models are commonly embedded within the (FEM) by inserting zero-thickness cohesive elements along potential crack paths or interfaces. Popular traction-separation laws include bilinear forms, which linearly increase to \sigma_c and then soften to zero at \delta_f, and exponential laws, which provide smoother softening tails to better capture gradual damage. The total fracture energy \Gamma dissipated in the cohesive zone is given by the \Gamma = \int_0^{\delta_f} \sigma(\delta) \, d\delta, which must be calibrated to match experimental values for accurate simulations. These implementations allow for progressive damage initiation based on a failure criterion, such as maximum or interaction. The advantages of cohesive zone modeling lie in its ability to naturally simulate initiation from concentrations without predefined paths, as well as with branching or kinking in response to mixed-mode loading. It has been extensively applied to composite materials, where it captures between plies, and to joints, predicting failure modes like peel or under varying geometries. Recent extensions of cohesive zone models incorporate rate dependence to address dynamic fracture scenarios, such as high-speed impacts, by introducing time-dependent parameters in the traction-separation law that increase peak strength and with loading . These viscous or viscoplastic enhancements enable modeling of strain--sensitive materials like polymers or metals under ballistic conditions.

Transition from Linear to Elastic-Plastic Regimes

The transition from linear fracture mechanics (LEFM) to elastic-plastic fracture mechanics (EPFM) occurs when the plastic zone at the crack tip becomes sufficiently large relative to the crack length or specimen dimensions, invalidating the small-scale yielding assumption of LEFM. A key quantitative criterion is the transition flaw size a_t = 2.5 \left( \frac{K_{Ic}}{\sigma_y} \right)^2 , where K_{Ic} is the and \sigma_y is the strength; LEFM remains valid for flaw sizes a \gg a_t, as the plastic zone size r_p is then much smaller than a. This threshold ensures that the stress field dominates, allowing accurate prediction of crack stability using stress intensity factors. In deformation theory, the choice between LEFM and EPFM is guided by the value relative to the limit J_{el}, defined as the J at the onset of significant in the load-displacement curve (typically 5-10% deviation from ). LEFM applies when J < J_{el}, where the response remains predominantly , while EPFM is required otherwise to account for contributions. Practical guidelines for assessing this include Anderson's criteria, such as r_p / a < 0.025 and r_p / b < 0.025 (where b is the uncracked ), alongside the loss-of- , which quantifies the deviation from behavior in experimental records. These ensure the zone from LEFM adjustments remains contained. Design implications emphasize flaw tolerance, particularly in pressure vessels, where 1980s standards incorporated fracture mechanics to evaluate allowable flaw sizes and prevent under internal pressure. For instance, these guidelines used transition criteria to set inspection thresholds, balancing safety with operational efficiency in petrochemical applications. Hybrid approaches, such as the two-parameter K-J method, bridge intermediate regimes by correlating the K with the , enabling consistent toughness assessment across elastic-to-plastic transitions without abrupt model switching.

Specialized Applications and Advanced Methods

Fracture in Concrete and Quasi-Brittle Materials

Fracture mechanics in concrete and quasi-brittle materials differs fundamentally from that in metals due to the heterogeneous microstructure, where aggregates and cement paste create a large fracture process zone involving microcracking and aggregate bridging across the crack surfaces. In metals, failure is typically ductile with localized plasticity at the crack tip, whereas concrete exhibits quasi-brittle behavior characterized by a tension-softening curve, denoted as σ(w), which describes the tensile stress σ as a function of crack opening width w, reflecting progressive damage and energy dissipation through aggregate interlocking and pull-out mechanisms. This softening leads to nonlinear fracture, contrasting with the linear elastic assumptions dominant in metallic fracture analysis. A seminal approach to modeling this behavior is Hillerborg's fictitious crack model, introduced in , which treats the fracture process zone as a cohesive where tensile stresses act across the fictitious surfaces according to a softening law σ(w) until complete separation. The model integrates mechanics with finite element analysis, defining the total energy G_f as the area under the σ(w) curve, typically ranging from 100 to 200 N/m for ordinary , which quantifies the energy required for propagation and accounts for the material's inherent toughness. This cohesive law captures the post-peak softening observed in tension tests, enabling predictions of growth in structures without relying on sharp tip singularities. To address the structural size effect in quasi-brittle , Bazant's size effect law, proposed in , describes how nominal strength σ_N decreases with increasing specimen size D due to the diminishing influence of boundary layers around the relative to the overall structure. The law is expressed as \sigma_N = \frac{B f_t}{\sqrt{1 + D/D_0}}, where f_t is the tensile strength, B is a geometry-dependent constant, and D_0 is a related to the and size, reflecting the transition from plastic-like behavior in small structures to dominance in large ones. This formulation arises from the release rate depending on the crack band width, approximately three times the maximum size, and has been validated through extensive testing on specimens. Experimental determination of fracture parameters like G_f follows RILEM recommendations, particularly using three-point tests on notched beams to measure the load-crack mouth opening curve and compute the work-of-fracture. In these tests, beams of standardized dimensions (e.g., span-to-depth ratio of 4) are loaded until complete , with G_f calculated as the total dissipated divided by the area, ensuring crack propagation and reliable quantification of softening behavior. Recent RILEM efforts, such as TC 187-SOC, extend this to directly obtaining the full σ(w) curve for enhanced model calibration. In applications, these models inform the design of beams by predicting widths and load-carrying capacity under tension or , where aggregate bridging and softening influence ultimate strength and serviceability. For seismic design, fracture mechanics aids in assessing and energy dissipation in beam-column joints, optimizing to control and prevent brittle during earthquakes. Recent advances in , incorporating steel or glass fibers at dosages of 0.5-2% by volume, significantly enhance G_f (up to 500-1000 N/m) and post- toughness by amplifying bridging effects, improving seismic performance and reducing in high-performance structures.

Atomistic and Discrete Fracture Mechanics

Atomistic fracture mechanics employs computational simulations at the nanoscale to elucidate the fundamental mechanisms governing crack initiation and propagation, bridging the gap between continuum theories and microscopic material behavior. (MD) simulations, utilizing empirical such as the embedded atom method (EAM) for metals, model atomic interactions to predict dynamics and bond breaking under stress. These approaches reveal how macroscopic stress intensity factors serve as boundary conditions to drive atomic-scale responses, enabling the study of in materials like FCC metals where plastic deformation via emission dominates brittle . A seminal contribution is the Rice-Thomson criterion, which posits that dislocation emission from a crack tip occurs when the stress intensity reaches a critical value that overcomes the resistance, favoring ductile behavior over brittle in crystals. This 1974 model incorporates Peierls stress, the frictional resistance arising from the atomic periodicity that impedes glide, providing a theoretical threshold for instability at the crack tip. At the atomic scale, the Griffith energy balance manifests as the competition between creation and elastic strain release, but discreteness effects like trapping can stabilize cracks below or above the continuum prediction, altering the effective energy. Discrete fracture models extend these ideas by representing materials as networks of interconnected elements, such as beam lattices with randomly assigned strengths to capture heterogeneity and damage localization. In these frameworks, fracture emerges from progressive bond failure, simulating quasi-brittle failure without relying on predefined crack paths. , introduced in 2000, reformulates in a nonlocal manner, where material points interact over a finite horizon, naturally accommodating discontinuities like cracks without special singularity treatments. This approach has been applied to model dynamic fracture in brittle solids, highlighting nonlocal damage evolution. MD simulations of fracture reveal intrinsic limits on crack propagation speeds, typically approaching but not exceeding the speed due to dynamic instabilities and energy dissipation at the atomic level. In amorphous materials like , crack advance induces localized amorphization and shear transformation zones, enhancing apparent toughness through non-affine deformations rather than discrete dislocations. Recent advances in the incorporate potentials, which train on quantum mechanical data to achieve accuracy at MD scales, enabling larger simulations of complex fracture paths in metals and ceramics. In semiconductors, quantum effects such as electron-hole pair generation during rapid crack motion introduce nonadiabatic influences, leading to quantized crack velocities and bursts that deviate from classical predictions.

Fatigue and Dynamic Fracture Considerations

Fatigue crack growth refers to the progressive extension of cracks under cyclic loading conditions, a critical phenomenon in structural integrity assessment. The Paris law provides a fundamental empirical relation for describing the crack growth rate in the stable propagation regime, expressed as \frac{da}{dN} = C (\Delta K)^m, where \frac{da}{dN} is the crack extension per cycle, \Delta K is the range, and C and m are material-specific constants typically determined experimentally. This law, originally derived from linear fracture mechanics principles, applies primarily to the mid-range of crack growth rates where environmental and microstructural effects are minimal. Below a stress intensity factor range, \Delta K_{th}, crack growth becomes negligible, often on the order of $10^{-8} to $10^{-10} m/cycle, preventing propagation in low-stress environments. Fatigue crack growth typically exhibits three regimes: Region I near the threshold, dominated by crack closure and environmental influences; Region II, the Paris regime with linear logarithmic dependence on \Delta K; and Region III, where growth accelerates toward instability as K_{max} approaches the K_c. Experimental measurement of fatigue crack growth rates commonly employs compact tension (CT) specimens, standardized under ASTM E647, which allow precise monitoring of crack length via compliance techniques or optical methods under constant amplitude loading. These tests reveal how factors like load ratio R = K_{min}/K_{max} influence \Delta K_{th}, with negative R values often increasing the threshold due to enhanced crack closure. Cyclic plasticity plays a central role in fatigue mechanisms, where repeated plastic deformation at the crack tip leads to irreversible slip, persistent slip bands, and microcrack coalescence, driving incremental advance per cycle. In applications such as wings, where components endure millions of load cycles from and pressurization, fracture mechanics-based predictions using the Paris law guide damage-tolerant design, ensuring safe operation until detected cracks are repaired. Dynamic fracture addresses crack propagation under high-speed or impact loading, where inertial and rate-dependent effects dominate. Crack speeds are theoretically limited to below the speed c_R, approximately 0.93 times the shear wave speed c_s in isotropic materials, beyond which wave reflections destabilize the tip. Freund's seminal linear solution for steady-state crack growth quantifies the release rate as G(v) = G_0 \frac{1 - v^2/c_d^2}{\sqrt{1 - v^2/c_s^2}}, where G(v) is the dynamic energy release rate, G_0 is the quasi-static value, v is the crack speed, and c_d and c_s are the dilatational and wave speeds, respectively; this highlights how G decreases with increasing v, requiring higher driving for sustained propagation. Energy dissipation in dynamic fracture often occurs through crack branching, where the primary deflects to create secondary paths, absorbing and limiting speeds to about 0.6–0.8 c_R in brittle materials. Viscous losses, arising from rate-sensitive material behavior or fluid interactions, further attenuate propagation by increasing effective under rapid loading. Dynamic fracture toughness is assessed via impact testing, such as precracked Charpy or modified three-point bend specimens under ASTM E1922, which capture rate-sensitive K_{Id} values often 2–4 times higher than static K_{Ic} due to adiabatic heating and viscoplasticity. These methods reveal enhanced resistance in metals like steels, where dynamic effects elevate toughness by up to 50% at speeds exceeding 10 m/s. Recent advancements in the 2020s have extended fatigue considerations to additively manufactured (AM) materials, where microstructural defects like porosity and lack-of-fusion influence \Delta K_{th} and Paris law parameters, often reducing fatigue life by 50–80% compared to wrought alloys without post-processing. Heat treatments and hot isostatic pressing mitigate these effects, improving growth resistance in AM titanium alloys for aerospace components. As of 2025, machine learning models integrated with fracture mechanics simulations have advanced defect-tolerant design for AM parts, enabling better prediction of fatigue crack growth in complex geometries.

References

  1. [1]
    [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 ...
  2. [2]
    VI. The phenomena of rupture and flow in solids - Journals
    The phenomena of rupture and flow in solids. Alan Arnold Griffith. Google Scholar · Find this author on PubMed · Search for more papers by this author. Alan ...
  3. [3]
    [PDF] 1 Overview: linear elastic fracture mechanics
    The field of fracture mechanics is concerned with the analytical description of the propagation of cracks. Classical elasticity theory predicts locally infinite ...<|control11|><|separator|>
  4. [4]
    [PDF] Engineering fracture mechanics has played a vital role in the ...
    This paper presents an overview of the computational, analytical and experimental strategies for fracture mechanics for fatigue, fracture and damage tolerance ...
  5. [5]
    History of fracture mechanics - Cornell University
    Who started all this fracture mechanics stuff? In 1913, a man named C.E. Inglis looked at a thin plate of glass (?) with an elliptical hole in the middle in ...
  6. [6]
    Griffith's Energy Release Rate - Fracture Mechanics
    Alan Arnold Griffith's energy-based analysis of cracks in 1920 is considered to be the birth of the field of fracture mechanics [1].
  7. [7]
    George R. Irwin | Memorial Tributes: Volume 10
    George observed that the fracture process in metals involved nonelastic work at the crack tip. This observation permitted him to modify the Griffith theory by ...
  8. [8]
    Fracture Toughness Testing - an overview | ScienceDirect Topics
    This committee developed the first widely recognized fracture toughness test standard, the plane strain fracture toughness method ASTM E 399-70T (1970) for the ...
  9. [9]
    Proceedings of the First International Conference on Fracture: Held ...
    Title, Proceedings of the First International Conference on Fracture: Held in Sendai, Japan, September 12-17, 1965, Volume 1
  10. [10]
    [PDF] The Elastic Stress Field around a Crack Tip - Elsevier
    Schematic of the basic fracture modes: (a) Mode I (opening), (b) Mode II (sliding),. (c) Mode III (tearing). 1. Mode I (Opening Mode): The two crack surfaces ...<|control11|><|separator|>
  11. [11]
    [PDF] Review of Fracture Toughness (G, K, J, CTOD, CTOA) Testing and ...
    As defined in ASTM E1823, the plane strain fracture toughness, KIc, is the crack-extension resistance under conditions of crack-tip plane strain in mode I for ...Missing: seminal | Show results with:seminal<|control11|><|separator|>
  12. [12]
    Breakdown of Continuum Fracture Mechanics at the Nanoscale
    Feb 26, 2015 · Based on Saint-Venant's principle, the plate height H is set to be large enough (H = 8W) to avoid undesirable effects from the loading points to ...
  13. [13]
    [PDF] A Brief History of the Crack Tip Stress Intensity Factor and
    Abstract The primary objective of this work is to discuss the origins, background and development of the elastic crack tip stress intensity factor, K, ...
  14. [14]
    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 ; Received: February 19, 1956 ; Published: September 1, 1957.
  15. [15]
    Energy Release Rate - an overview | ScienceDirect Topics
    The energy release rate \( G \) is defined as the decrease of the total potential energy per increase in fracture length or surface area, representing the ...
  16. [16]
    [PDF] Fracture mechanics lecture 2
    The crack will grow if the energy release rate equals the fracture energy: G = Γ. The energy release rate is the driving force for the extension of the crack.
  17. [17]
    Energy Concepts for Crack Growth - Engineering Library
    The energy release rate is sometimes referred to as the rate of strain energy flux flowing toward a crack tip as the crack extends. Let us consider the load ...
  18. [18]
    Fracture Mechanics | SpringerLink
    Aug 13, 2012 · Providing the stress and strain fields near the tip of a crack, linear elastic fracture mechanics (LEFM) enables to predict when this crack will ...Missing: G_I = | Show results with:G_I =
  19. [19]
    Plastic Yielding at Crack Tip | Engineering Library
    3.1 Irwin's Model. To determine the plastic zone at the crack tip, Irwin presented a simple model assuming the material is elastic-perfectly plastic. Consider ...
  20. [20]
    [PDF] SSC-345 ELASTIC - PLASTIC FRACTURE MECHANICS
    Irwin (11) proposed a simple plastic zone correction to . the stress intensity factor. An alternative plastic zone correction was developed by. Dugdale (12) ...
  21. [21]
    [PDF] Stress-intensity factors
    May 19, 2020 · • Stress ahead of crack tip cannot exceed yield stress ... • The energy criterion was the original concept of fracture mechanics: KI =.
  22. [22]
    The contributions of George Irwin to elastic-plastic fracture ...
    This paper describes a simple analysis based on elasticity theory and fracture mechanics that will predict characteristics of shallow hydraulic fractures that ...
  23. [23]
    (PDF) Fracture Mechanics - ResearchGate
    Aug 24, 2015 · framework of linear fracture mechanics. Fig. 2.22. Effect of plasticity in reducing the stresses around the tip of a crack. We will also discuss ...
  24. [24]
    [PDF] Some Observations on Damage Tolerance Analyses in Pressure ...
    and limitations of, LEFM. II. Pressure Vessel Configurations. Figure 1 shows a cylindrical metallic PV and two COPVs. COPVs have a metallic liner and a ...
  25. [25]
    Fracture Toughness of Metal Castings - IntechOpen
    Fracture toughness of a material may be defined as the critical value of the stress intensity factor, the latter depending on a combination of the stress at ...
  26. [26]
    On the validity of LEFM methods to investigate the fracture behavior ...
    Mar 1, 2019 · ... plastic zone ahead of crack tip, can be interpreted as a nonlinear ... specimen thickness and L is the total length. Fig. 1. Download ...
  27. [27]
    Fracture Mechanics Fundamentals | Engineering Library
    Several hundred liberty ships fractured extensively during World War II. The failures occurred primarily because of the change from riveted to welded ...
  28. [28]
    [PDF] Fracture Toughness Testing: G, K, J, CTOD, CTOA
    The first fracture toughness test standard ASTM E399 [13] was developed to determine the point value of plane strain fracture toughness at or near to the onset ...
  29. [29]
    CTOD and pipelines – the past, present and future - TWI Global
    Originally developed from research at TWI in the UK during the 1960's, CTOD was an ideal parameter for characterising the fracture toughness of medium strength ...<|separator|>
  30. [30]
    A Path Independent Integral and the Approximate Analysis of Strain ...
    Rice, J. R. (June 1, 1968). "A Path Independent Integral and the ... Applicability of J Estimation Procedures for Overmatched SE(T) Fracture Specimens.
  31. [31]
    Application of J Integral for the Fracture Assessment of Welded ...
    Elastic–plastic fracture mechanics parameters can also be used for structural damage tolerance assessment, fitness-for-service evaluation, residual strength ...
  32. [32]
    [PDF] NASA TM X-7192 FRACTURE MECHANICS
    Jan 31, 1974 · This size is used to adjust the rate fracture occurs. of crack propagation during successive stress cycles according to the relative sizes of ...
  33. [33]
    R-Curves - an overview | ScienceDirect Topics
    The R-curve is defined as a representation of the increasing fracture resistance of a material as a function of crack length, quantifying the length-dependent ...
  34. [34]
    Crack blunting, crack bridging and resistance-curve fracture ...
    Crack bridging by uncracked ligaments, observed directly by microscopy and X-ray tomography, was identified as a major toughening mechanism, with further ...Missing: stretching | Show results with:stretching
  35. [35]
    Estimations on J-integral and tearing modulus T from a single ...
    The objective of this paper is to investigate the implications of the load separation criterion for evaluating ductile fracture mechanics parameters. This ...Missing: σ_y^ dJ/ da seminal
  36. [36]
    [PDF] Investigaiton of Tearing Instability Phenomena in ASTM Steel. - DTIC
    The material tearing modulus was calculated from the expression,. Tmaterial. (E/Oo )(dJ/da). (3) where. E = elastic modulus. Go= flow stress (a + Uu)/2. dJ/da ...Missing: σ_y^ seminal paper
  37. [37]
    E561 Standard Test Method for KR Curve Determination - ASTM
    May 9, 2023 · This test method covers the determination of the resistance to fracture of metallic materials under Mode I loading at static rates.Missing: metals | Show results with:metals
  38. [38]
    (PDF) Experimental Estimation of JR Curves from Load-CMOD ...
    Apr 1, 2016 · This paper demonstrates the possibility of experimental and numerical determination of fracture mechanics behavior of characteristic ...Missing: seminal | Show results with:seminal
  39. [39]
    J-integral resistance curve testing and evaluation - Academia.edu
    In this paper a critical review is presented of the history and current state of the art of J-integral resistance curve testing and experimental evaluation ...<|control11|><|separator|>
  40. [40]
    [PDF] Cohesive Zone Models: A Critical Review of Traction-Separation ...
    The present paper critically reviews traction-separation relationships of cohesive fracture with an emphasis on potential-based constitu- tive models. There are ...
  41. [41]
    An improved description of the exponential Xu and Needleman ...
    Numerical aspects of cohesive-zone models ... This paper checks the consistency of some published exponential and bilinear mixed-mode cohesive zone models.
  42. [42]
    Application of Cohesive Zone Modeling to Composite Bonded Repairs
    Aug 7, 2025 · It was verified that CZM are able to predict with accuracy the damage mechanisms and strength of composite bonded repairs, thus constituting a ...
  43. [43]
    A cohesive zone element for mode I modelling of adhesives ...
    Citation Excerpt : The major advantage of their proposed formulation is its complete independency of the cohesive zone length.
  44. [44]
    A generalized rate-dependent cohesive zone model with energy ...
    Oct 16, 2025 · This paper establishes a novel rate-dependent cohesive zone model (CZM) to characterize the rate-dependent damage behavior of interfaces.
  45. [45]
    Improved cohesive zone model: integrating strain rate, plastic strain ...
    Jul 6, 2023 · In this paper, artificial compliance and spurious oscillation in bilinear cohesive law are investigated through dynamic simulations.
  46. [46]
  47. [47]
    FRACTURE MECHANICS F o u rt h E d i t i o n Fundamentals and ...
    ... book on fracture mechanics. However, dimensional analysis is an impor- tant tool for developing mathematical models of physical phenomena, and it can help ...
  48. [48]
    [PDF] Two-Parameter Fracture Mechanics Theory and Applications - DTIC
    The two parameters, J and Q, have distinct roles: J sets the size scale of the process zone over which large stresses and strains develop, while Q scales the ...
  49. [49]
    Fracture resistance on aggregate bridging crack in concrete
    It is found that G I-bridging will firstly increase with the development of crack and then stay constant once the initial crack tip opening displacement reaches ...
  50. [50]
    A new method for determining the tension-softening curve of concrete
    The simplified bilinear tension-softening curve of concrete can be used to analyze the fracture behavior of concrete structures.
  51. [51]
    [PDF] State-of-the-art review on fracture analysis of concrete structural ...
    This review covers fracture models, tension softening models, crack growth analysis, and remaining life prediction for concrete structural components.
  52. [52]
    Analysis of crack formation and crack growth in concrete by means ...
    Abstract. A method is presented in which fracture mechanics is introduced into finite element analysis by means of a model where stresses are assumed to act ...
  53. [53]
    analysis of crack formation and crack growth in concrete by means ...
    The method integrates fracture mechanics with finite element analysis (FEM) for crack growth evaluation. It explains the difference between bending strength and ...
  54. [54]
    [PDF] SIZE EFFECT IN BLUNT FRACTURE: CONCRETE, ROCK, METAL ...
    Study of these deviations is the objective of this paper. Understanding of the structural size effect is important for correct interpretation of test data. Most ...
  55. [55]
    On the measurement of concrete fracture energy using three-point ...
    RILEM Draft Recommendation: 'Determination of fracture energy of mortar and concrete by means of three-point bend tests on notched beams',Mater. Struct. 18 ( ...
  56. [56]
    187-SOC : Experimental determination of the stress-crack opening ...
    From their work, four RILEM recommendations issued: FMC 1-Determination of the fracture energy of mortar and concrete by means of three-point bend tests on ...
  57. [57]
    Fracture mechanics modeling of reinforced concrete joints ...
    The developed fracture mechanics modeling was applicable for identifying propagation of a crack in concrete structural members such as beam-column connections.
  58. [58]
    Fiber-reinforced cementitious composites: recent advances and ...
    Apr 1, 2025 · This paper comprehensively reviews fiber-reinforced concrete (FRC), focusing on workability, fiber-matrix interaction, mechanical properties ...
  59. [59]
    A review of lattice type model in fracture mechanics
    The lattice model is a discrete model that is typically used to simulate the fracture process of brittle materials. This review summarizes the main ...
  60. [60]
    Beam network model for fracture of materials with hierarchical ...
    Jan 19, 2021 · We introduce a beam network model for hierarchically patterned materials. In these materials, load-parallel gaps intercept stress transmission in the load ...
  61. [61]
    Reformulation of elasticity theory for discontinuities and long-range ...
    As an aid to the modeling of such materials, a new framework for the basic equations of continuum mechanics, called the 'peridynamic' formulation, is proposed.
  62. [62]
    Dynamic Instability in Intergranular Fracture | Phys. Rev. Lett.
    Jun 30, 2005 · In homogeneous media at speeds greater than 1 / 3 of the Rayleigh wave speed c R , these instabilities result in crack branching and roughness ...
  63. [63]
  64. [64]
    Atomistic measurement and modeling of intrinsic fracture toughness ...
    Nov 4, 2022 · We present advances in experimentation and machine learning–inspired algorithms for accurate interatomic potential ... machine learning, atomistic ...