Energy release rate (fracture mechanics)

Last updated

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, [1] [2] 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.

Contents

Definition

Plot of Load vs. Displacement Plot of Load vs. Load-Displacement.png
Plot of Load vs. Displacement

The energy release rate is defined [3] as the instantaneous loss of total potential energy per unit crack growth area ,

where the total potential energy is written in terms of the total strain energy , surface traction , displacement , and body force by

The first integral is over the surface of the material, and the second is over its volume .

The figure on the right shows the plot of an external force vs. the load-point displacement , in which the area under the curve is the strain energy. The white area between the curve and the -axis is referred to as the complementary energy. In the case of a linearly-elastic material, is a straight line and the strain energy is equal to the complementary energy.

Prescribed displacement

In the case of prescribed displacement, the strain energy can be expressed in terms of the specified displacement and the crack surface , and the change in this strain energy is only affected by the change in fracture surface area: . Correspondingly, the energy release rate in this case is expressed as [3]

Here is where one can accurately refer to as the strain energy release rate.

Prescribed loads

When the load is prescribed instead of the displacement, the strain energy needs to be modified as . The energy release rate is then computed as [3]

If the material is linearly-elastic, then and one may instead write

G in two-dimensional cases

In the cases of two-dimensional problems, the change in crack growth area is simply the change in crack length times the thickness of the specimen. Namely, . Therefore, the equation for computing can be modified for the 2D case:

One can refer to the example calculations embedded in the next section for further information. Sometimes, the strain energy is written using , an energy-per-unit thickness. This gives

Relation to stress intensity factors

The energy release rate is directly related to the stress intensity factor associated with a given two-dimensional loading mode (Mode-I, Mode-II, or Mode-III) when the crack grows straight ahead. [3] This is applicable to cracks under plane stress, plane strain, and antiplane shear.

For Mode-I, the energy release rate rate is related to the Mode-I stress intensity factor for a linearly-elastic material by

where is related to Young's modulus and Poisson's ratio depending on whether the material is under plane stress or plane strain:

For Mode-II, the energy release rate is similarly written as

For Mode-III (antiplane shear), the energy release rate now is a function of the shear modulus ,

For an arbitrary combination of all loading modes, these linear elastic solutions may be superposed as

Relation to fracture toughness

Crack growth is initiated when the energy release rate overcomes a critical value , which is a material property,

Under Mode-I loading, the critical energy release rate is then related to the Mode-I fracture toughness , another material property, by

Calculating G

There are a variety of methods available for calculating the energy release rate given material properties, specimen geometry, and loading conditions. Some are dependent on certain criteria being satisfied, such as the material being entirely elastic or even linearly-elastic, and/or that the crack must grow straight ahead. The only method presented that works arbitrarily is that using the total potential energy. If two methods are both applicable, they should yield identical energy release rates.

Total potential energy

The only method to calculate for arbitrary conditions is to calculate the total potential energy and differentiate it with respect to the crack surface area. This is typically done by:

all in terms of the crack surface area.

Compliance method

If the material is linearly elastic, the computation of its energy release rate can be much simplified. In this case, the Load vs. Load-point Displacement curve is linear with a positive slope, and the displacement per unit force applied is defined as the compliance, [3]

The corresponding strain energy (area under the curve) is equal to [3]

Using the compliance method, one can show that the energy release rate for both cases of prescribed load and displacement come out to be [3]

Multiple specimen methods for nonlinear materials

Graphical Illustration of G under Fixed Displacement and Fixed Load Conditions Graphical Illustration of G under Fixed Displacement and Fixed Load Conditions.png
Graphical Illustration of G under Fixed Displacement and Fixed Load Conditions

In the case of prescribed displacement, holding the crack length fixed, the energy release rate can be computed by [3]

while in the case of prescribed load, [3]

As one can see, in both cases, the energy release rate times the change in surface returns the area between curves, which indicates the energy dissipated for the new surface area as illustrated in the right figure [3]

Crack closure integral

Since the energy release rate is defined as the negative derivative of the total potential energy with respect to crack surface growth, the energy release rate may be written as the difference between the potential energy before and after the crack grows. After some careful derivation, this leads one to the crack closure integral [3]

where is the new fracture surface area, are the components of the traction released on the top fracture surface as the crack grows, are the components of the crack opening displacement (the difference in displacement increments between the top and bottom crack surfaces), and the integral is over the surface of the material .

The crack closure integral is valid only for elastic materials but is still valid for cracks that grow in any direction. Nevertheless, for a two-dimensional crack that does indeed grow straight ahead, the crack closure integral simplifies to [3]

where is the new crack length, and the displacement components are written as a function of the polar coordinates and .

J-integral

In certain situations, the energy release rate can be calculated using the J-integral, i.e. , using [3]

where is the elastic strain energy density, is the component of the unit vector normal to , the curve used for the line integral, are the components of the traction vector , where is the stress tensor, and are the components of the displacement vector.

This integral is zero over a simple closed path and is path independent, allowing any simple path starting and ending on the crack faces to be used to calculate . In order to equate the energy release rate to the J-integral, , the following conditions must be met:

The J-integral may be calculated with these conditions violated, but then . When they are not violated, one can then relate the energy release rate and the J-integral to the elastic moduli and the stress intensity factors using [3]

Computational methods in fracture mechanics

A handful of methods exist for calculating with finite elements. Although a direct calculation of the J-integral is possible (using the strains and stresses outputted by FEA), approximate approaches for some type of crack growth exist and provide reasonable accuracy with straightforward calculations. This section will elaborate on some relatively simple methods for fracture analysis utilizing numerical simulations.

Nodal release method

If the crack is growing straight, the energy release rate can be decomposed as a sum of 3 terms associated with the energy in each 3 modes. As a result, the Nodal Release method (NR) can be used to determine from FEA results. The energy release rate is calculated at the nodes of the finite element mesh for the crack at an initial length and extended by a small distance . First, we calculate the displacement variation at the node of interest (before and after the crack tip node is released). Secondly, we keep track of the nodal force outputted by FEA. Finally, we can find each components of using the following formulas:

Figure 1: Crack debounding between two consecutive time steps using nodal release (NR) Nodal Release Debounded.png
Figure 1: Crack debounding between two consecutive time steps using nodal release (NR)

Where is the width of the element bounding the crack tip. The accuracy of the method highly depends on the mesh refinement, both because the displacement and forces depend on it, and because . Note that the equations above are derived using the crack closure integral.

If the energy release rate exceeds a critical value, the crack will grow. In this case, a new FEA simulation is performed (for the next time step) where the node at the crack tip is released. For a bounded substrate, we may simply stop enforcing fixed Dirichlet boundary conditions at the crack tip node of the previous time step (i.e. displacements are no longer restrained). For a symmetric crack, we would need to update the geometry of the domain with a longer crack opening (and therefore generate a new mesh [5] ).

Modified crack closure integral

Similar to the Nodal Release Method, the Modified Crack Closure Integral (MCCI) is a method for calculating the energy release rate utilizing FEA nodal displacements and forces . [6] [7] Where represents the direction corresponding to the Cartesian basis vectors with origin at the crack tip, and represents the nodal index. MCCI is more computationally efficient than the nodal release method because it only requires one analysis for each increment of crack growth.

A necessary condition for the MCCI method is uniform element length along the crack face in the direction. Additionally, this method requires sufficient discretization such that over the length of one element stress fields are self-similar. This implies that as the crack propagates. Below are examples of the MCCI method with two types of common finite elements.

4-node elements

Figure 2: Schematic of the MCCI process for 4-node linear rectangular elements. 4-node.png
Figure 2: Schematic of the MCCI process for 4-node linear rectangular elements.

The 4-node square linear elements seen in Figure 2 have a distance between nodes and equal to Consider a crack with its tip located at node Similar to the nodal release method, if the crack were to propagate one element length along the line of symmetry (parallel to the -axis) the crack opening displacement would be the displacement at the previous crack tip, i.e. and the force at the new crack tip would be Since the crack growth is assumed to be self-similar the displacement at node after the crack propagates is equal to the displacement at node before the crack propagates. This same concept can be applied to the forces at node and Utilizing the same method shown in the nodal release section we recover the following equations for energy release rate:

Where (displacement above and below the crack face respectively). Because we have a line of symmetry parallel to the crack, we can assume

Thus,

8-node elements

Figure 3: Schematic of the MCCI process for 9-node quadratic rectangular elements. 8-node.png
Figure 3: Schematic of the MCCI process for 9-node quadratic rectangular elements.

The 8-node rectangular elements seen in Figure 3 have quadratic basis functions. The process for calculating G is the same as the 4-node elements with the exception that (the crack growth over one element) is now the distance from node to Once again, making the assumption of self-similar straight crack growth the energy release rate can be calculated utilizing the following equations:

Like with the nodal release method the accuracy of MCCI is highly dependent on the level of discretization along the crack tip, i.e. Accuracy also depends on element choice. A mesh of 8-node quadratic elements can produce more accurate results than a mesh of 4-node linear elements with the same number of degrees of freedom [8] in the mesh.

Domain integral approach for J

Figure 4: Contour domain integral for the J-integral Contour Domain Integral J-Integral.png
Figure 4: Contour domain integral for the J-integral

The J-integral may be calculated directly using the finite element mesh and shape functions. [9] We consider a domain contour as shown in figure 4 and choose an arbitrary smooth function such that on and on .

For linear elastic cracks growing straight ahead, . The energy release rate can then be calculated over the area bounded by the contour using an updated formulation:

The formula above may be applied to any annular area surrounding the crack tip (in particular, a set of neighboring elements can be used). This method is very accurate, even with a coarse mesh around the crack tip (one may choose an integration domain located far away, with stresses and displacement less sensitive to mesh refinement)

2-D crack tip singular elements

The above-mentioned methods for calculating energy release rate asymptotically approach the actual solution with increased discretization but fail to fully capture the crack tip singularity. More accurate simulations can be performed by utilizing quarter-point elements around the crack tip. [10] These elements have a built-in singularity which more accurately produces stress fields around the crack tip. The advantage of the quarter-point method is that it allows for coarser finite element meshes and greatly reduces computational cost. Furthermore, these elements are derived from small modifications to common finite elements without requiring special computational programs for analysis. For the purposes of this section elastic materials will be examined, although this method can be extended to elastic-plastic fracture mechanics. [11] [12] [13] [14] Assuming perfect elasticity the stress fields will experience a crack tip singularity.

8-node isoparametric element

Figure 5: Mapped and Parent Element of the 8-node isoparametric finite element. Note the quarter point location of the nodes in the mapped element. Finalfig5.png
Figure 5: Mapped and Parent Element of the 8-node isoparametric finite element. Note the quarter point location of the nodes in the mapped element.

The 8-node quadratic element is described by Figure 5 in both parent space with local coordinates and and by the mapped element in physical/global space by and The parent element is mapped from the local space to the physical space by the shape functions and the degree of freedom coordinates The crack tip is located at or

In a similar way, displacements (defined as ) can also be mapped.

A property of shape functions in the finite element method is compact support, specifically the Kronecker delta property (i.e. at node and zero at all other nodes). This results in the following shape functions for the 8-node quadratic elements: [8]

When considering a line in front of the crack that is co-linear with the - axis (i.e. ) all basis functions are zero except for

Calculating the normal strain involves using the chain rule to take the derivative of displacement with respect to

Figure 6: Triangular Element mapped from 8-node rectangular element. Triangle256.png
Figure 6: Triangular Element mapped from 8-node rectangular element.

If the nodes are spaced evenly on the rectangular element then the strain will not contain the singularity. By moving nodes 5 and 8 position to a quarter of the length of the element closer to the crack tip as seen in figure 5, the mapping from becomes:

Solving for and taking the derivative results in:

Plugging this result into the equation for strain the final result is obtained:

By moving the mid-nodes to a quarter position results in the correct crack tip singularity.

Other element types

Figure 7: Natural Triangle Element Triagneles2.png
Figure 7: Natural Triangle Element

The rectangular element method does not allow for singular elements to be easily meshed around the crack tip. This impedes the ability to capture the angular dependence of the stress fields which is critical in determining the crack path. Also, except along the element edges the singularity exists in a very small region near the crack tip. Figure 6 shows another quarter-point method for modeling this singularity. The 8-node rectangular element can be mapped into a triangle. [15] This is done by collapsing the nodes on the line to the mid-node location and shifting the mid-nodes on to the quarter-point location. The collapsed rectangle can more easily surround the crack tip but requires that the element edges be straight or the accuracy of calculating the stress intensity factor will be reduced.

A better candidate for the quarter-point method is the natural triangle as seen in Figure 7. The element's geometry allows for the crack tip to be easily surrounded and meshing is simplified. Following the same procedure described above, the displacement and strain field for the triangular elements are:

This method reproduces the first two terms of the Williams solutions [16] with a constant and singular term.

An advantage of the quarter-point method is that it can be easily generalized to 3-dimensional models. This can greatly reduce computation when compared to other 3-dimensional methods but can lead to errors if that crack tip propagates with a large degree of curvature. [17]

See also

Related Research Articles

<span class="mw-page-title-main">Dirac delta function</span> Generalized function whose value is zero everywhere except at zero

In mathematical analysis, the Dirac delta function, also known as the unit impulse, is a generalized function on the real numbers, whose value is zero everywhere except at zero, and whose integral over the entire real line is equal to one. Since there is no function having this property, to model the delta "function" rigorously involves the use of limits or, as is common in mathematics, measure theory and the theory of distributions.

In physics, a Langevin equation is a stochastic differential equation describing how a system evolves when subjected to a combination of deterministic and fluctuating ("random") forces. The dependent variables in a Langevin equation typically are collective (macroscopic) variables changing only slowly in comparison to the other (microscopic) variables of the system. The fast (microscopic) variables are responsible for the stochastic nature of the Langevin equation. One application is to Brownian motion, which models the fluctuating motion of a small particle in a fluid.

In special relativity, four-momentum (also called momentum–energy or momenergy) is the generalization of the classical three-dimensional momentum to four-dimensional spacetime. Momentum is a vector in three dimensions; similarly four-momentum is a four-vector in spacetime. The contravariant four-momentum of a particle with relativistic energy E and three-momentum p = (px, py, pz) = γmv, where v is the particle's three-velocity and γ the Lorentz factor, is

In mathematics, the Hodge star operator or Hodge star is a linear map defined on the exterior algebra of a finite-dimensional oriented vector space endowed with a nondegenerate symmetric bilinear form. Applying the operator to an element of the algebra produces the Hodge dual of the element. This map was introduced by W. V. D. Hodge.

A directional derivative is a concept in multivariable calculus that measures the rate at which a function changes in a particular direction at a given point.

In Hamiltonian mechanics, a canonical transformation is a change of canonical coordinates (q, p) → that preserves the form of Hamilton's equations. This is sometimes known as form invariance. Although Hamilton's equations are preserved, it need not preserve the explicit form of the Hamiltonian itself. Canonical transformations are useful in their own right, and also form the basis for the Hamilton–Jacobi equations and Liouville's theorem.

Second-order linear partial differential equations (PDEs) are classified as either elliptic, hyperbolic, or parabolic. Any second-order linear PDE in two variables can be written in the form

In physics, the Hamilton–Jacobi equation, named after William Rowan Hamilton and Carl Gustav Jacob Jacobi, is an alternative formulation of classical mechanics, equivalent to other formulations such as Newton's laws of motion, Lagrangian mechanics and Hamiltonian mechanics.

<span class="mw-page-title-main">Jacobi integral</span>

In celestial mechanics, Jacobi's integral is the only known conserved quantity for the circular restricted three-body problem. Unlike in the two-body problem, the energy and momentum of the system are not conserved separately and a general analytical solution is not possible. The integral has been used to derive numerous solutions in special cases.

In the theory of general relativity, linearized gravity is the application of perturbation theory to the metric tensor that describes the geometry of spacetime. As a consequence, linearized gravity is an effective method for modeling the effects of gravity when the gravitational field is weak. The usage of linearized gravity is integral to the study of gravitational waves and weak-field gravitational lensing.

In mathematics, in particular in algebraic geometry and differential geometry, Dolbeault cohomology (named after Pierre Dolbeault) is an analog of de Rham cohomology for complex manifolds. Let M be a complex manifold. Then the Dolbeault cohomology groups depend on a pair of integers p and q and are realized as a subquotient of the space of complex differential forms of degree (p,q).

In physics and fluid mechanics, a Blasius boundary layer describes the steady two-dimensional laminar boundary layer that forms on a semi-infinite plate which is held parallel to a constant unidirectional flow. Falkner and Skan later generalized Blasius' solution to wedge flow, i.e. flows in which the plate is not parallel to the flow.

<span class="mw-page-title-main">Peridynamics</span>

Peridynamics is a non-local formulation of continuum mechanics that is oriented toward deformations with discontinuities, especially fractures. Originally, bond-based peridynamic has been introduced, wherein, internal interactions forces between a material point and all the other ones with which it can interact, are modeled as a central forces field. This type of forces field can be imagined as a mesh of bonds connecting each point of the body with every other interacting points within a certain distance which depends on material property, called peridynamic horizon. Later, to overcome bond-based framework limitations for the material Poisson’s ratio, state-base peridynamics, has been formulated. Its characteristic feature is that the force exchanged between a point and another one is influenced by the deformation state of all other bonds relative to its interaction zone.

In mathematics, the Malgrange–Ehrenpreis theorem states that every non-zero linear differential operator with constant coefficients has a Green's function. It was first proved independently by Leon Ehrenpreis and Bernard Malgrange (1955–1956).

In mathematics, the spectral theory of ordinary differential equations is the part of spectral theory concerned with the determination of the spectrum and eigenfunction expansion associated with a linear ordinary differential equation. In his dissertation, Hermann Weyl generalized the classical Sturm–Liouville theory on a finite closed interval to second order differential operators with singularities at the endpoints of the interval, possibly semi-infinite or infinite. Unlike the classical case, the spectrum may no longer consist of just a countable set of eigenvalues, but may also contain a continuous part. In this case the eigenfunction expansion involves an integral over the continuous part with respect to a spectral measure, given by the Titchmarsh–Kodaira formula. The theory was put in its final simplified form for singular differential equations of even degree by Kodaira and others, using von Neumann's spectral theorem. It has had important applications in quantum mechanics, operator theory and harmonic analysis on semisimple Lie groups.

<span class="mw-page-title-main">Cnoidal wave</span> Nonlinear and exact periodic wave solution of the Korteweg–de Vries equation

In fluid dynamics, a cnoidal wave is a nonlinear and exact periodic wave solution of the Korteweg–de Vries equation. These solutions are in terms of the Jacobi elliptic function cn, which is why they are coined cnoidal waves. They are used to describe surface gravity waves of fairly long wavelength, as compared to the water depth.

In mathematical physics, the Whitham equation is a non-local model for non-linear dispersive waves.

<span class="mw-page-title-main">Symmetry in quantum mechanics</span> Properties underlying modern physics

Symmetries in quantum mechanics describe features of spacetime and particles which are unchanged under some transformation, in the context of quantum mechanics, relativistic quantum mechanics and quantum field theory, and with applications in the mathematical formulation of the standard model and condensed matter physics. In general, symmetry in physics, invariance, and conservation laws, are fundamentally important constraints for formulating physical theories and models. In practice, they are powerful methods for solving problems and predicting what can happen. While conservation laws do not always give the answer to the problem directly, they form the correct constraints and the first steps to solving a multitude of problems.

In combustion, Frank-Kamenetskii theory explains the thermal explosion of a homogeneous mixture of reactants, kept inside a closed vessel with constant temperature walls. It is named after a Russian scientist David A. Frank-Kamenetskii, who along with Nikolay Semenov developed the theory in the 1930s.

References

  1. Li, F.Z.; Shih, C.F.; Needleman, A. (1985). "A comparison of methods for calculating energy release rates". Engineering Fracture Mechanics. 21 (2): 405–421. doi:10.1016/0013-7944(85)90029-3. ISSN   0013-7944.
  2. Rice, J.R.; Budiansky, B. (1973). "Conservation laws and energy-release rates". Journal of Applied Mechanics. 40 (1): 201–3. Bibcode:1973JAM....40..201B. doi:10.1115/1.3422926. S2CID   13910502.
  3. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 Alan Zehnder (2012). Fracture Mechanics. London ; New York : Springer Science+Business Media. ISBN   9789400725942.
  4. Soboyejo, W. O. (2003). "11.6.5 Equivalence of G and K". Mechanical properties of engineered materials. Marcel Dekker. ISBN   0-8247-8900-8. OCLC 300921090.
  5. Tradegard, A. (1998-07-15). "FEM-remeshing technique applied to crack growth problems". Computer Methods in Applied Mechanics and Engineering. 160 (1–2): 115–131. Bibcode:1998CMAME.160..115T. doi:10.1016/s0045-7825(97)00287-9.
  6. Rybicki, E.F.; Kanninen, M.F. (January 1977). "A finite element calculation of stress intensity factors by a modified crack closure integral". Engineering Fracture Mechanics. 9 (4): 931–938. doi:10.1016/0013-7944(77)90013-3. ISSN   0013-7944.
  7. Sethuraman, R.; Maiti, S.K. (January 1988). "Finite element based computation of strain energy release rate by modified crack closure integral". Engineering Fracture Mechanics. 30 (2): 227–231. doi:10.1016/0013-7944(88)90226-3. ISSN   0013-7944.
  8. 1 2 Zehnder, Alan T. (2012-01-03). Fracture mechanics. Dordrecht. ISBN   9789400725959. OCLC   773034407.{{cite book}}: CS1 maint: location missing publisher (link)
  9. Zehnder, Alan T. (2012). Fracture Mechanics. Lecture Notes in Applied and Computational Mechanics. Vol. 62. Dordrecht: Springer Netherlands. doi:10.1007/978-94-007-2595-9. ISBN   9789400725942.
  10. Henshell, R. D.; Shaw, K. G. (1975). "Crack tip finite elements are unnecessary". International Journal for Numerical Methods in Engineering. 9 (3): 495–507. Bibcode:1975IJNME...9..495H. doi:10.1002/nme.1620090302. ISSN   0029-5981.
  11. Barsoum, Roshdy S. (1977). "Triangular quarter-point elements as elastic and perfectly-plastic crack tip elements". International Journal for Numerical Methods in Engineering. 11 (1): 85–98. Bibcode:1977IJNME..11...85B. doi:10.1002/nme.1620110109. ISSN   0029-5981.
  12. Sun, C.T.; Jin, Z.-H. (2012), "Elastic-Plastic Fracture Criteria", Fracture Mechanics, Elsevier, pp. 171–187, doi:10.1016/b978-0-12-385001-0.00007-9, ISBN   9780123850010
  13. Stern, Morris (1979). "Families of consistent conforming elements with singular derivative fields". International Journal for Numerical Methods in Engineering. 14 (3): 409–421. Bibcode:1979IJNME..14..409S. doi:10.1002/nme.1620140307. ISSN   0029-5981.
  14. Levy, N.; Marcal, P.V.; Ostergren, W.J.; Rice, J.R. (June 1971). "Small scale yielding near a crack in plane strain: A finite element analysis". International Journal of Fracture Mechanics. 7 (2): 143–156. doi:10.1007/bf00183802. ISSN   0020-7268. S2CID   11088286.
  15. Barsoum, Roshdy S. (1976). "On the use of isoparametric finite elements in linear fracture mechanics". International Journal for Numerical Methods in Engineering. 10 (1): 25–37. Bibcode:1976IJNME..10...25B. doi:10.1002/nme.1620100103. ISSN   0029-5981.
  16. Williams, M.L (1959). "The stresses around a fault or crack in dissimilar media" (PDF). Bulletin of the Seismological Society of America. 49 (2): 199–204. Bibcode:1959BuSSA..49..199W. doi:10.1785/BSSA0490020199.
  17. Peano, A.; Pasini, A. (February 1982). "A warning against misuse of quarter-point elements". International Journal for Numerical Methods in Engineering. 18 (2): 314–320. Bibcode:1982IJNME..18..314P. doi:10.1002/nme.1620180212. ISSN   0029-5981.