Damage mechanics

Last updated

Damage mechanics is concerned with the representation, or modeling, of damage of materials that is suitable for making engineering predictions about the initiation, propagation, and fracture of materials without resorting to a microscopic description that would be too complex for practical engineering analysis. [1]

Contents

Damage mechanics illustrates the typical engineering approach to model complex phenomena. To quote Dusan Krajcinovic, "It is often argued that the ultimate task of engineering research is to provide not so much a better insight into the examined phenomenon but to supply a rational predictive tool applicable in design". [2] Damage mechanics is a topic of applied mechanics that relies heavily on continuum mechanics. Most of the work on damage mechanics uses state variables to represent the effects of damage on the stiffness and remaining life of the material that is damaging as a result of thermomechanical load and ageing. [3] The state variables may be measurable, e.g., crack density, or inferred from the effect they have on some macroscopic property, such as stiffness, coefficient of thermal expansion, remaining life, etc. The state variables have conjugate thermodynamic forces that motivate further damage. Initially the material is pristine, or intact. A damage activation criterion is needed to predict damage initiation. Damage evolution does not progress spontaneously after initiation, thus requiring a damage evolution model. In plasticity like formulations, the damage evolution is controlled by a hardening function but this requires additional phenomenological parameters that must be found through experimentation, which is expensive, time consuming, and virtually no one does. On the other hand, micromechanics of damage formulations are able to predict both damage initiation and evolution without additional material properties. [4]

Creep Continuum Damage Mechanics

When mechanical structures are exposed to temperatures exceeding one-third of the melting temperature of the material of construction, time-dependent deformation (creep) and associated material degradation mechanisms become dominant modes of structural failure. While these deformation and damage mechanisms originate at the microscale where discrete processes dominate, practical application of failure theories to macroscale components is most readily achieved using the formalism of continuum mechanics. In this context, microscopic damage is idealized as a continuous state variable defined at all points within a structure. State equations are defined which govern the time evolution of damage. These equations may be readily integrated into finite element codes to analyze the damage evolution in complex 3D structures and calculate how long a component may safely be used before failure occurs.

Lumped damage state variable

L. M. Kachanov [5] and Y. N. Rabotnov [6] suggested the following evolution equations for the creep strain ε and a lumped damage state variable ω:

Where, is the creep strain rate, is the creep-rate multiplier, is the applied stress, is the creep stress exponent of the material of interest, is the rate of damage accumulation, is the damage-rate multiplier, and is the damage stress exponent.

In this simple case, the strain rate is governed by power-law creep with the stress enhanced by the damage state variable as damage accumulates. The damage term ω is interpreted as a distributed loss of load bearing area which results in an increased local stress at the microscale. The time to failure is determined by integrating the damage evolution equation from an initial undamaged state to a specified critical damage . If is taken to be 1, this results in the following prediction for a structure loaded under a constant uniaxial stress :

Model parameters and n are found by fitting the creep strain rate equation at zero damage to minimum creep rate measurements. Model parameters and m are found by fitting the above equation to creep rupture life data.

Mechanistically informed damage state variables

While easy to apply, the lumped damage model proposed by Kachanov [7] and Robotnov [8] is limited by the fact that the damage state variable cannot be directly tied to a specific mechanism of strain and damage evolution. Correspondingly, extrapolation of the model beyond the original dataset of test data is not justified. This limitation was remedied by researchers such as A.C.F. Cocks, [9] M.F. Ashby, [10] and B.F. Dyson, [11] who proposed mechanistically informed strain and damage evolution equations. Extrapolation using such equations is justified if the dominant damage mechanism remains the same at the conditions of interest.

Void-growth by Power-Law Creep

In the power-law creep regime, global deformation is controlled by glide and climb of dislocations. If internal voids are present within the microstructure, global structural continuity requires that the voids must both elongate and expand laterally, further reducing the local section. When cast in the damage mechanics formalism, the growth of internal voids by power-law creep can be represented by the following equations. [12] [13]

Where, is the creep-rate multiplier, is the applied stress, n is the creep stress exponent, is the average initial void radius, and d is the grain size.

Void-growth by Boundary Diffusion

At very high temperature and/or low stresses, void growth on grain boundaries is primarily controlled by the diffusive flux of vacancies along the grain boundary. As matter diffuses away from the void and plates onto the adjacent grain boundaries, a roughly spherical void is maintained by rapid diffusion of vacancies along the surface of the void. When cast in the damage mechanics formalism, the growth of internal voids by boundary diffusion can be represented by the following equations. [14] [15]

Where, is the creep-rate multiplier, is the applied stress, is the center-to-center void spacing, is the grain size, is the grain-boundary diffusion coefficient, is the grain boundary thickness, is the atomic volume, is Boltzmann’s constant, and is the absolute temperatures. It is noted that factors present in are very similar to the Coble creep pre-factors due to the similarity of the two mechanisms.

Precipitate Coarsening

Many modern steels and alloys are designed such that precipitates will precipitate either within the matrix or along grain boundaries during casting. These precipitates restrict dislocation motion and, if present on grain boundaries, grain boundary sliding during creep. Many precipitates are not thermodynamically stable and grow via diffusion when exposed to elevated temperatures. As the precipitates coarsen, their ability to restrict dislocation motion decreases as the average spacing between particles increases, thus decreasing the required Orowan stress for bowing. In the case of grain boundary precipitates, precipitate growth means that fewer grain boundaries are impeded from grain boundary sliding. When cast into the damage mechanics formalism, precipitation coarsening and its effect on strain rate may be represented by the following equations. [16]

Where, is the creep-rate multiplier, is the applied stress, is the creep-rate stress exponent, is a parameter linking the precipitation damage to the strain rate, determines the rate of precipitate coarsening.

Combining Damage Mechanisms

Multiple damage mechanism can be combined to represent a broader range of phenomena. For instance, if both void-growth by power-law creep and precipitate coarsening are relevant mechanisms, the following combined set of equations may be used:

Note that both damage mechanisms are included in the creep strain rate equation. The precipitate coarsening damage mechanisms influences the void-growth damage mechanism as the void-growth mechanism depends on the global strain rate. The precipitate growth mechanisms is only time and temperature dependent and hence does not depend on the void-growth damage .

Multiaxial Effects

The preceding equations are valid under uniaxial tension only. When a multiaxial state of stress is present in the system, each equation must be adapted so that the driving multiaxial stress is considered. For void-growth by power-law creep, the relevant stress is the von Mises stress as this drives the global creep deformation; however, for void-growth by boundary diffusion, the maximum principal stress drives the vacancy flux.

See also

Related Research Articles

Dynamic mechanical analysis is a technique used to study and characterize materials. It is most useful for studying the viscoelastic behavior of polymers. A sinusoidal stress is applied and the strain in the material is measured, allowing one to determine the complex modulus. The temperature of the sample or the frequency of the stress are often varied, leading to variations in the complex modulus; this approach can be used to locate the glass transition temperature of the material, as well as to identify transitions corresponding to other molecular motions.

A Maxwell material is the most simple model viscoelastic material showing properties of a typical liquid. It shows viscous flow on the long timescale, but additional elastic resistance to fast deformations. It is named for James Clerk Maxwell who proposed the model in 1867. It is also known as a Maxwell fluid.

<span class="mw-page-title-main">Creep (deformation)</span> Tendency of a solid material to move slowly or deform permanently under mechanical stress

In materials science, creep is the tendency of a solid material to undergo slow deformation while subject to persistent mechanical stresses. It can occur as a result of long-term exposure to high levels of stress that are still below the yield strength of the material. Creep is more severe in materials that are subjected to heat for long periods and generally increase as they near their melting point.

In materials science and continuum mechanics, viscoelasticity is the property of materials that exhibit both viscous and elastic characteristics when undergoing deformation. Viscous materials, like water, resist shear flow and strain linearly with time when a stress is applied. Elastic materials strain when stretched and immediately return to their original state once the stress is removed.

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.

In mechanics, virtual work arises in the application of the principle of least action to the study of forces and movement of a mechanical system. The work of a force acting on a particle as it moves along a displacement is different for different displacements. Among all the possible displacements that a particle may follow, called virtual displacements, one will minimize the action. This displacement is therefore the displacement followed by the particle according to the principle of least action.

The work of a force on a particle along a virtual displacement is known as the virtual work.

<span class="mw-page-title-main">Two-state quantum system</span> Simple quantum mechanical system

In quantum mechanics, a two-state system is a quantum system that can exist in any quantum superposition of two independent quantum states. The Hilbert space describing such a system is two-dimensional. Therefore, a complete basis spanning the space will consist of two independent states. Any two-state system can also be seen as a qubit.

A Kelvin-Voigt material, also called a Voigt material, is the most simple model viscoelastic material showing typical rubbery properties. It is purely elastic on long timescales, but shows additional resistance to fast deformation. It is named after the British physicist and engineer Lord Kelvin and German physicist Woldemar Voigt.

Betti's theorem, also known as Maxwell–Betti reciprocal work theorem, discovered by Enrico Betti in 1872, states that for a linear elastic structure subject to two sets of forces {Pi} i=1,...,n and {Qj}, j=1,2,...,n, the work done by the set P through the displacements produced by the set Q is equal to the work done by the set Q through the displacements produced by the set P. This theorem has applications in structural engineering where it is used to define influence lines and derive the boundary element method.

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

Coble creep, a form of diffusion creep, is a mechanism for deformation of crystalline solids. Contrasted with other diffusional creep mechanisms, Coble creep is similar to Nabarro–Herring creep in that it is dominant at lower stress levels and higher temperatures than creep mechanisms utilizing dislocation glide. Coble creep occurs through the diffusion of atoms in a material along grain boundaries. This mechanism is observed in polycrystals or along the surface in a single crystal, which produces a net flow of material and a sliding of the grain boundaries.

In geology, a deformation mechanism is a process occurring at a microscopic scale that is responsible for changes in a material's internal structure, shape and volume. The process involves planar discontinuity and/or displacement of atoms from their original position within a crystal lattice structure. These small changes are preserved in various microstructures of materials such as rocks, metals and plastics, and can be studied in depth using optical or digital microscopy.

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

Microvoid coalescence (MVC) is a high energy microscopic fracture mechanism observed in the majority of metallic alloys and in some engineering plastics.

The Larson–Miller relation, also widely known as the Larson–Miller parameter and often abbreviated LMP, is a parametric relation used to extrapolate experimental data on creep and rupture life of engineering materials.

<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.

Thermo-mechanical fatigue is the overlay of a cyclical mechanical loading, that leads to fatigue of a material, with a cyclical thermal loading. Thermo-mechanical fatigue is an important point that needs to be considered, when constructing turbine engines or gas turbines.

Dislocation creep is a deformation mechanism in crystalline materials. Dislocation creep involves the movement of dislocations through the crystal lattice of the material, in contrast to diffusion creep, in which diffusion is the dominant creep mechanism. It causes plastic deformation of the individual crystals, and thus the material itself.

<span class="mw-page-title-main">Grain boundary sliding</span> Material deformation mechanism

Grain boundary sliding (GBS) is a material deformation mechanism where grains slide against each other. This occurs in polycrystalline material under external stress at high homologous temperature and low strain rate and is intertwined with creep. Homologous temperature describes the operating temperature relative to the melting temperature of the material. There are mainly two types of grain boundary sliding: Rachinger sliding, and Lifshitz sliding. Grain boundary sliding usually occurs as a combination of both types of sliding. Boundary shape often determines the rate and extent of grain boundary sliding.

<span class="mw-page-title-main">Rock mass plasticity</span>

Plasticity theory for rocks is concerned with the response of rocks to loads beyond the elastic limit. Historically, conventional wisdom has it that rock is brittle and fails by fracture while plasticity is identified with ductile materials. In field scale rock masses, structural discontinuities exist in the rock indicating that failure has taken place. Since the rock has not fallen apart, contrary to expectation of brittle behavior, clearly elasticity theory is not the last word.

Nabarro–Herring creep is a mode of deformation of crystalline materials that occurs at low stresses and held at elevated temperatures in fine-grained materials. In Nabarro–Herring creep, atoms diffuse through the crystals, and the creep rate varies inversely with the square of the grain size so fine-grained materials creep faster than coarser-grained ones. NH creep is solely controlled by diffusional mass transport. This type of creep results from the diffusion of vacancies from regions of high chemical potential at grain boundaries subjected to normal tensile stresses to regions of lower chemical potential where the average tensile stresses across the grain boundaries are zero. Self-diffusion within the grains of a polycrystalline solid can cause the solid to yield to an applied shearing stress, the yielding being caused by a diffusional flow of matter within each crystal grain away from boundaries where there is a normal pressure and toward those where there is a normal tension. Atoms migrating in the opposite direction account for the creep strain. The creep strain rate is derived in the next section. NH creep is more important in ceramics than metals as dislocation motion is more difficult to effect in ceramics.

Anelasticity is a property of materials that describes their behaviour when undergoing deformation. Its formal definition does not include the physical or atomistic mechanisms but still interprets the anelastic behaviour as a manifestation of internal relaxation processes. It is a behaviour differing from elastic behaviour.

References

  1. Krajcinovic, D., Damage mechanics (1989) Mechanics of Materials, 8 (2-3), pp. 117-197.
  2. Dusan Krajcinovic, Mechanics of Materials 8 (1989) 169.
  3. Struik, L C E, Physical aging in amorphous polymers and other materials, Elsevier Scientific Pub. Co.; New York, 1978, ISBN   9780444416551.
  4. Barbero, E.J., Cortes, D.H., A mechanistic model for transverse damage initiation, evolution, and stiffness reduction in laminated composites (2010) Composites Part B: Engineering, 41 (2), pp. 124-132.
  5. Kachanov, Lazar M. (1 April 1999). "Rupture Time Under Creep Conditions". International Journal of Fracture. 97 (1): 11–18. doi:10.1023/A:1018671022008. S2CID   116979654.
  6. Rabotnov, Y. N. (1969). "Creep rupture". Applied Mechanics: 342–349. doi:10.1007/978-3-642-85640-2_26. ISBN   978-3-642-85642-6.
  7. Kachanov, Lazar M. (1 April 1999). "Rupture Time Under Creep Conditions". International Journal of Fracture. 97 (1): 11–18. doi:10.1023/A:1018671022008. S2CID   116979654.
  8. Rabotnov, Y. N. (1969). "Creep rupture". Applied Mechanics: 342–349. doi:10.1007/978-3-642-85640-2_26. ISBN   978-3-642-85642-6.
  9. Cocks, A. C. F.; Ashby, M. F. (1 January 1982). "On creep fracture by void growth". Progress in Materials Science. 27 (3): 189–244. doi:10.1016/0079-6425(82)90001-9.
  10. Cocks, A. C. F.; Ashby, M. F. (1 January 1982). "On creep fracture by void growth". Progress in Materials Science. 27 (3): 189–244. doi:10.1016/0079-6425(82)90001-9.
  11. Dyson, B.F. (1988). "Creep and fracture of metals : mechanisms and mechanics" (PDF). Revue de Physique Appliquée. 23 (4): 605–613. doi:10.1051/rphysap:01988002304060500.
  12. Cocks, A. C. F.; Ashby, M. F. (1 January 1982). "On creep fracture by void growth". Progress in Materials Science. 27 (3): 189–244. doi:10.1016/0079-6425(82)90001-9.
  13. Dyson, B.F. (1988). "Creep and fracture of metals : mechanisms and mechanics" (PDF). Revue de Physique Appliquée. 23 (4): 605–613. doi:10.1051/rphysap:01988002304060500.
  14. Cocks, A. C. F.; Ashby, M. F. (1 January 1982). "On creep fracture by void growth". Progress in Materials Science. 27 (3): 189–244. doi:10.1016/0079-6425(82)90001-9.
  15. Dyson, B.F. (1988). "Creep and fracture of metals : mechanisms and mechanics" (PDF). Revue de Physique Appliquée. 23 (4): 605–613. doi:10.1051/rphysap:01988002304060500.
  16. Dyson, B. F. (1992). "Materials Data Requirements, Creep Damage Mechanisms, and Predictive Models". High Temperature Structural Design.