Stress intensity factor

Last updated
Polar coordinates at the crack tip. Cracktipcoords.svg
Polar coordinates at the crack tip.

In fracture mechanics, the stress intensity factor (K) is used to predict the stress state ("stress intensity") near the tip of a crack or notch caused by a remote load or residual stresses. [1] It is a theoretical construct usually applied to a homogeneous, linear elastic material and is useful for providing a failure criterion for brittle materials, and is a critical technique in the discipline of damage tolerance. The concept can also be applied to materials that exhibit small-scale yielding at a crack tip.

Contents

The magnitude of K depends on specimen geometry, the size and location of the crack or notch, and the magnitude and the distribution of loads on the material. It can be written as: [2] [3]

where is a specimen geometry dependent function of the crack length, a, and the specimen width, W, and σ is the applied stress.

Linear elastic theory predicts that the stress distribution () near the crack tip, in polar coordinates () with origin at the crack tip, has the form [4]

where K is the stress intensity factor (with units of stress × length1/2) and is a dimensionless quantity that varies with the load and geometry. Theoretically, as r goes to 0, the stress goes to resulting in a stress singularity. [5] Practically however, this relation breaks down very close to the tip (small r) because plasticity typically occurs at stresses exceeding the material's yield strength and the linear elastic solution is no longer applicable. Nonetheless, if the crack-tip plastic zone is small in comparison to the crack length, the asymptotic stress distribution near the crack tip is still applicable.

Stress intensity factors for various modes

Mode I, Mode II, and Mode III crack loading. Fracture modes v2.svg
Mode I, Mode II, and Mode III crack loading.

In 1957, G. Irwin found that the stresses around a crack could be expressed in terms of a scaling factor called the stress intensity factor. He found that a crack subjected to any arbitrary loading could be resolved into three types of linearly independent cracking modes. [6] These load types are categorized as Mode I, II, or III as shown in the figure. Mode I is an opening (tensile) mode where the crack surfaces move directly apart. Mode II is a sliding (in-plane shear) mode where the crack surfaces slide over one another in a direction perpendicular to the leading edge of the crack. Mode III is a tearing (antiplane shear) mode where the crack surfaces move relative to one another and parallel to the leading edge of the crack. Mode I is the most common load type encountered in engineering design.

Different subscripts are used to designate the stress intensity factor for the three different modes. The stress intensity factor for mode I is designated and applied to the crack opening mode. The mode II stress intensity factor, , applies to the crack sliding mode and the mode III stress intensity factor, , applies to the tearing mode. These factors are formally defined as: [7]

Relationship to energy release rate and J-integral

In plane stress conditions, the strain energy release rate () for a crack under pure mode I, or pure mode II loading is related to the stress intensity factor by:

where is the Young's modulus and is the Poisson's ratio of the material. The material is assumed to be an isotropic, homogeneous, and linear elastic. The crack has been assumed to extend along the direction of the initial crack

For plane strain conditions, the equivalent relation is a little more complicated:

For pure mode III loading,

where is the shear modulus. For general loading in plane strain, the linear combination holds:

A similar relation is obtained for plane stress by adding the contributions for the three modes.

The above relations can also be used to connect the J-integral to the stress intensity factor because

Critical stress intensity factor

The stress intensity factor, , is a parameter that amplifies the magnitude of the applied stress that includes the geometrical parameter (load type). Stress intensity in any mode situation is directly proportional to the applied load on the material. If a very sharp crack, or a V-notch can be made in a material, the minimum value of can be empirically determined, which is the critical value of stress intensity required to propagate the crack. This critical value determined for mode I loading in plane strain is referred to as the critical fracture toughness () of the material. has units of stress times the root of a distance (e.g. MN/m3/2). The units of imply that the fracture stress of the material must be reached over some critical distance in order for to be reached and crack propagation to occur. The Mode I critical stress intensity factor, , is the most often used engineering design parameter in fracture mechanics and hence must be understood if we are to design fracture tolerant materials used in bridges, buildings, aircraft, or even bells.

Polishing cannot detect a crack. Typically, if a crack can be seen it is very close to the critical stress state predicted by the stress intensity factor[ citation needed ].

G–criterion

The G-criterion is a fracture criterion that relates the critical stress intensity factor (or fracture toughness) to the stress intensity factors for the three modes. This failure criterion is written as [8]

where is the fracture toughness, for plane strain and for plane stress. The critical stress intensity factor for plane stress is often written as .


Examples

Infinite plate: Uniform uniaxial stress

The stress intensity factor for an assumed straight crack of length perpendicular to the loading direction, in an infinite plane, having a uniform stress field is [5] [7]

Crack in an infinite plate under mode I loading. CrackInfinitePlate.svg
Crack in an infinite plate under mode I loading.

Penny-shaped crack in an infinite domain

The stress intensity factor at the tip of a penny-shaped crack of radius in an infinite domain under uniaxial tension is [1]

Penny-shaped crack in an infinite domain under uniaxial tension. PennyShapedCrack.svg
Penny-shaped crack in an infinite domain under uniaxial tension.

Finite plate: Uniform uniaxial stress

If the crack is located centrally in a finite plate of width and height , an approximate relation for the stress intensity factor is [7]

If the crack is not located centrally along the width, i.e., , the stress intensity factor at location A can be approximated by the series expansion [7] [9]

where the factors can be found from fits to stress intensity curves [7] :6 for various values of . A similar (but not identical) expression can be found for tip B of the crack. Alternative expressions for the stress intensity factors at A and B are [10] :175

where

with

In the above expressions is the distance from the center of the crack to the boundary closest to point A. Note that when the above expressions do not simplify into the approximate expression for a centered crack.

Crack in a finite plate under mode I loading. CrackFinitePlate.svg
Crack in a finite plate under mode I loading.

Edge crack in a plate under uniaxial stress

For a plate having dimensions containing an unconstrained edge crack of length , if the dimensions of the plate are such that and , the stress intensity factor at the crack tip under a uniaxial stress is [5]

For the situation where and , the stress intensity factor can be approximated by

Edge crack in a finite plate under uniaxial stress. CrackEdgeFinitePlate.svg
Edge crack in a finite plate under uniaxial stress.

Infinite plate: Slanted crack in a biaxial stress field

For a slanted crack of length in a biaxial stress field with stress in the -direction and in the -direction, the stress intensity factors are [7] [11]

where is the angle made by the crack with the -axis.

A slanted crack in a thin plate under biaxial load. CrackSlantedPlateBiaxialLoad.svg
A slanted crack in a thin plate under biaxial load.

Crack in a plate under point in-plane force

Consider a plate with dimensions containing a crack of length . A point force with components and is applied at the point () of the plate.

For the situation where the plate is large compared to the size of the crack and the location of the force is relatively close to the crack, i.e., , , , , the plate can be considered infinite. In that case, for the stress intensity factors for at crack tip B () are [11] [12]

where

with , , for plane strain, for plane stress, and is the Poisson's ratio. The stress intensity factors for at tip B are

The stress intensity factors at the tip A () can be determined from the above relations. For the load at location ,

Similarly for the load ,

A crack in a plate under the action of a localized force with components
F
x
{\displaystyle F_{x}}
and
F
y
{\displaystyle F_{y}}
. CrackFinitePlatePointForce.svg
A crack in a plate under the action of a localized force with components and .

Loaded crack in a plate

If the crack is loaded by a point force located at and , the stress intensity factors at point B are [7]

If the force is distributed uniformly between , then the stress intensity factor at tip B is

A loaded crack in a plate. LoadedCrackPlate.svg
A loaded crack in a plate.

Compact tension specimen

The stress intensity factor at the crack tip of a compact tension specimen is [13]

where is the applied load, is the thickness of the specimen, is the crack length, and is the width of the specimen.

Compact tension specimen for fracture toughness testing. CompactTensionSpecimen.svg
Compact tension specimen for fracture toughness testing.

Single-edge notch-bending specimen

The stress intensity factor at the crack tip of a single-edge notch-bending specimen is [13]

where is the applied load, is the thickness of the specimen, is the crack length, and is the width of the specimen.

Single-edge notch-bending specimen (also called three-point bending specimen) for fracture toughness testing SingleEdgeNotchBending.svg
Single-edge notch-bending specimen (also called three-point bending specimen) for fracture toughness testing

See also

Related Research Articles

<span class="mw-page-title-main">Fracture</span> Split of materials or structures under stress

Fracture is the separation of an object or material into two or more pieces under the action of stress. The fracture of a solid usually occurs due to the development of certain displacement discontinuity surfaces within the solid. If a displacement develops perpendicular to the surface, it is called a normal tensile crack or simply a crack; if a displacement develops tangentially, it is called a shear crack, slip band or dislocation.

<span class="mw-page-title-main">Buckling</span> Sudden change in shape of a structural component under load

In structural engineering, buckling is the sudden change in shape (deformation) of a structural component under load, such as the bowing of a column under compression or the wrinkling of a plate under shear. If a structure is subjected to a gradually increasing load, when the load reaches a critical level, a member may suddenly change shape and the structure and component is said to have buckled. Euler's critical load and Johnson's parabolic formula are used to determine the buckling stress of a column.

<span class="mw-page-title-main">Fracture mechanics</span> Field of mechanics that studies the propagation of cracks in materials

Fracture mechanics is the field of mechanics concerned with the study of the propagation of cracks in materials. It uses methods of analytical solid mechanics to calculate the driving force on a crack and those of experimental solid mechanics to characterize the material's resistance to fracture.

Thermal shock is a phenomenon characterized by a rapid change in temperature that results in a transient mechanical load on an object. The load is caused by the differential expansion of different parts of the object due to the temperature change. This differential expansion can be understood in terms of strain, rather than stress. When the strain exceeds the tensile strength of the material, it can cause cracks to form and eventually lead to structural failure.

<span class="mw-page-title-main">Fracture toughness</span> Stress intensity factor at which a cracks propagation increases drastically

In materials science, fracture toughness is the critical stress intensity factor of a sharp crack where propagation of the crack suddenly becomes rapid and unlimited. A component's thickness affects the constraint conditions at the tip of a crack with thin components having plane stress conditions and thick components having plane strain conditions. Plane strain conditions give the lowest fracture toughness value which is a material property. The critical value of stress intensity factor in mode I loading measured under plane strain conditions is known as the plane strain fracture toughness, denoted . When a test fails to meet the thickness and other test requirements that are in place to ensure plane strain conditions, the fracture toughness value produced is given the designation . Fracture toughness is a quantitative way of expressing a material's resistance to crack propagation and standard values for a given material are generally available.

<span class="mw-page-title-main">Three-point flexural test</span> Standard procedure for measuring modulus of elasticity in bending

The three-point bending flexural test provides values for the modulus of elasticity in bending , flexural stress , flexural strain and the flexural stress–strain response of the material. This test is performed on a universal testing machine with a three-point or four-point bend fixture. The main advantage of a three-point flexural test is the ease of the specimen preparation and testing. However, this method has also some disadvantages: the results of the testing method are sensitive to specimen and loading geometry and strain rate.

The J-integral represents a way to calculate the strain energy release rate, or work (energy) per unit fracture surface area, in a material. The theoretical concept of J-integral was developed in 1967 by G. P. Cherepanov and independently in 1968 by James R. Rice, who showed that an energetic contour path integral was independent of the path around a crack.

<span class="mw-page-title-main">Paris' law</span>

Paris' law is a crack growth equation that gives the rate of growth of a fatigue crack. The stress intensity factor characterises the load around a crack tip and the rate of crack growth is experimentally shown to be a function of the range of stress intensity seen in a loading cycle. The Paris equation is

<span class="mw-page-title-main">Fracture (geology)</span> Geologic discontinuity feature, often a joint or fault

A fracture is any separation in a geologic formation, such as a joint or a fault that divides the rock into two or more pieces. A fracture will sometimes form a deep fissure or crevice in the rock. Fractures are commonly caused by stress exceeding the rock strength, causing the rock to lose cohesion along its weakest plane. Fractures can provide permeability for fluid movement, such as water or hydrocarbons. Highly fractured rocks can make good aquifers or hydrocarbon reservoirs, since they may possess both significant permeability and fracture porosity.

<span class="mw-page-title-main">Viscoplasticity</span> Theory in continuum mechanics

Viscoplasticity is a theory in continuum mechanics that describes the rate-dependent inelastic behavior of solids. Rate-dependence in this context means that the deformation of the material depends on the rate at which loads are applied. The inelastic behavior that is the subject of viscoplasticity is plastic deformation which means that the material undergoes unrecoverable deformations when a load level is reached. Rate-dependent plasticity is important for transient plasticity calculations. The main difference between rate-independent plastic and viscoplastic material models is that the latter exhibit not only permanent deformations after the application of loads but continue to undergo a creep flow as a function of time under the influence of the applied load.

Material failure theory is an interdisciplinary field of materials science and solid mechanics which attempts to predict the conditions under which solid materials fail under the action of external loads. The failure of a material is usually classified into brittle failure (fracture) or ductile failure (yield). Depending on the conditions most materials can fail in a brittle or ductile manner or both. However, for most practical situations, a material may be classified as either brittle or ductile.

In fracture mechanics, the energy release rate, , is the rate at which energy is transformed as a material undergoes fracture. Mathematically, the energy release rate is expressed as the decrease in total potential energy per increase in fracture surface area, and is thus expressed in terms of energy per unit area. Various energy balances can be constructed relating the energy released during fracture to the energy of the resulting new surface, as well as other dissipative processes such as plasticity and heat generation. The energy release rate is central to the field of fracture mechanics when solving problems and estimating material properties related to fracture and fatigue.

Polymer fracture is the study of the fracture surface of an already failed material to determine the method of crack formation and extension in polymers both fiber reinforced and otherwise. Failure in polymer components can occur at relatively low stress levels, far below the tensile strength because of four major reasons: long term stress or creep rupture, cyclic stresses or fatigue, the presence of structural flaws and stress-cracking agents. Formations of submicroscopic cracks in polymers under load have been studied by x ray scattering techniques and the main regularities of crack formation under different loading conditions have been analyzed. The low strength of polymers compared to theoretically predicted values are mainly due to the many microscopic imperfections found in the material. These defects namely dislocations, crystalline boundaries, amorphous interlayers and block structure can all lead to the non-uniform distribution of mechanical stress.

<span class="mw-page-title-main">Cohesive zone model</span>

The cohesive zone model (CZM) is a model in fracture mechanics where fracture formation is regarded as a gradual phenomenon and separation of the crack surfaces takes place across an extended crack tip, or cohesive zone, and is resisted by cohesive tractions. The origin of this model can be traced back to the early sixties by Dugdale (1960) and Barenblatt (1962) to represent nonlinear processes located at the front of a pre-existent crack.

Crack closure is a phenomenon in fatigue loading, where the opposing faces of a crack remain in contact even with an external load acting on the material. As the load is increased, a critical value will be reached at which time the crack becomes open. Crack closure occurs from the presence of material propping open the crack faces and can arise from many sources including plastic deformation or phase transformation during crack propagation, corrosion of crack surfaces, presence of fluids in the crack, or roughness at cracked surfaces.

<span class="mw-page-title-main">Crack tip opening displacement</span>

Crack tip opening displacement (CTOD) or is the distance between the opposite faces of a crack tip at the 90° intercept position. The position behind the crack tip at which the distance is measured is arbitrary but commonly used is the point where two 45° lines, starting at the crack tip, intersect the crack faces. The parameter is used in fracture mechanics to characterize the loading on a crack and can be related to other crack tip loading parameters such as the stress intensity factor and the elastic-plastic J-integral.

In materials science, toughening refers to the process of making a material more resistant to the propagation of cracks. When a crack propagates, the associated irreversible work in different materials classes is different. Thus, the most effective toughening mechanisms differ among different materials classes. The crack tip plasticity is important in toughening of metals and long-chain polymers. Ceramics have limited crack tip plasticity and primarily rely on different toughening mechanisms.

The fracture of soft materials involves large deformations and crack blunting before propagation of the crack can occur. Consequently, the stress field close to the crack tip is significantly different from the traditional formulation encountered in the Linear elastic fracture mechanics. Therefore, fracture analysis for these applications requires a special attention. The Linear Elastic Fracture Mechanics (LEFM) and K-field are based on the assumption of infinitesimal deformation, and as a result are not suitable to describe the fracture of soft materials. However, LEFM general approach can be applied to understand the basics of fracture on soft materials. The solution for the deformation and crack stress field in soft materials considers large deformation and is derived from the finite strain elastostatics framework and hyperelastic material models.

<span class="mw-page-title-main">Crack growth equation</span>

A crack growth equation is used for calculating the size of a fatigue crack growing from cyclic loads. The growth of fatigue cracks can result in catastrophic failure, particularly in the case of aircraft. A crack growth equation can be used to ensure safety, both in the design phase and during operation, by predicting the size of cracks. In critical structure, loads can be recorded and used to predict the size of cracks to ensure maintenance or retirement occurs prior to any of the cracks failing.

Fastran is a computer program for calculating the rate of fatigue crack growth by combining crack growth equations and a simulation of the plasticity at the crack tip.

References

  1. 1 2 Anderson, T. L. (2005). Fracture mechanics: fundamentals and applications. CRC Press.
  2. Soboyejo, W. O. (2003). "11.6.2 Crack Driving Force and Concept of Similitude". Mechanical properties of engineered materials. Marcel Dekker. ISBN   0-8247-8900-8. OCLC   300921090.
  3. Janssen, M. (Michael) (2004). Fracture mechanics. Zuidema, J. (Jan), Wanhill, R. J. H. (2nd ed.). London: Spon Press. p. 41. ISBN   0-203-59686-2. OCLC   57491375.
  4. Hiroshi Tada; P. C. Paris; George R. Irwin (February 2000). The Stress Analysis of Cracks Handbook (3rd ed.). American Society of Mechanical Engineers.
  5. 1 2 3 Liu, M.; et al. (2015). "An improved semi-analytical solution for stress at round-tip notches" (PDF). Engineering Fracture Mechanics. 149: 134–143.
  6. 1 2 Suresh, S. (2004). Fatigue of Materials. Cambridge University Press. ISBN   978-0-521-57046-6.
  7. 1 2 3 4 5 6 7 Rooke, D. P.; Cartwright, D. J. (1976). Compendium of stress intensity factors. HMSO Ministry of Defence. Procurement Executive.
  8. Sih, G. C.; Macdonald, B. (1974), "Fracture mechanics applied to engineering problems-strain energy density fracture criterion", Engineering Fracture Mechanics, 6 (2): 361–386, doi:10.1016/0013-7944(74)90033-2
  9. Isida, M., 1966, Stress intensity factors for the tension of an eccentrically cracked strip, Transactions of the ASME Applied Mechanics Section, v. 88, p.94.
  10. Kathiresan, K.; Brussat, T. R.; Hsu, T. M. (1984). Advanced life analysis methods. Crack Growth Analysis Methods for Attachment Lugs. Flight Dynamics Laboratory, Air Force Wright Aeronautical Laboratories, AFSC W-P Air Force Base, Ohio.
  11. 1 2 Sih, G. C.; Paris, P. C. & Erdogan, F. (1962), "Crack-tip stress intensity factors for the plane extension and plate bending problem", Journal of Applied Mechanics, 29: 306–312, Bibcode:1962JAM....29..306S, doi:10.1115/1.3640546
  12. Erdogan, F. (1962), "On the stress distribution in plates with collinear cuts under arbitrary loads", Proceedings of the Fourth US National Congress of Applied Mechanics, 1: 547–574
  13. 1 2 Bower, A. F. (2009). Applied mechanics of solids. CRC Press.