Eigenstrain

Last updated

In continuum mechanics an eigenstrain is any mechanical deformation in a material that is not caused by an external mechanical stress, with thermal expansion often given as a familiar example. The term was coined in the 1970s by Toshio Mura, who worked extensively on generalizing their mathematical treatment. [1] A non-uniform distribution of eigenstrains in a material (e.g., in a composite material) leads to corresponding eigenstresses, which affect the mechanical properties of the material. [2]

Contents

Overview

Many distinct physical causes for eigenstrains exist, such as crystallographic defects, thermal expansion, the inclusion of additional phases in a material, and previous plastic strains. [3] All of these result from internal material characteristics, not from the application of an external mechanical load. As such, eigenstrains have also been referred to as “stress-free strains” [4] and “inherent strains”. [5] When one region of material experiences a different eigenstrain than its surroundings, the restraining effect of the surroundings leads to a stress state on both regions. [6] Analyzing the distribution of this residual stress for a known eigenstrain distribution or inferring the total eigenstrain distribution from a partial data set are both two broad goals of eigenstrain theory.

Analysis of eigenstrains and eigenstresses

Eigenstrain analysis usually relies on the assumption of linear elasticity, such that different contributions to the total strain are additive. In this case, the total strain of a material is divided into the elastic strain e and the inelastic eigenstrain :

where and indicate the directional components in 3 dimensions in Einstein notation.

Another assumption of linear elasticity is that the stress can be linearly related to the elastic strain and the stiffness by Hooke’s Law: [3]

In this form, the eigenstrain is not in the equation for stress, hence the term "stress-free strain". However, a non-uniform distribution of eigenstrain alone will cause elastic strains to form in response, and therefore a corresponding elastic stress. When performing these calculations, closed-form expressions for (and thus, the total stress and strain fields) can only be found for specific geometries of the distribution of . [5]

Ellipsoidal inclusion in an infinite medium

Ellipsoidal eigenstrain inclusion Ellipsoidal eigenstrain inclusion.svg
Ellipsoidal eigenstrain inclusion

One of the earliest examples providing such a closed-form solution analyzed a ellipsoidal inclusion of material with a uniform eigenstrain, constrained by an infinite medium with the same elastic properties. [6] This can be imagined with the figure on the right. The inner ellipse represents the region . The outer region represents the extent of if it fully expanded to the eigenstrain without being constrained by the surrounding . Because the total strain, shown by the solid outlined ellipse, is the sum of the elastic and eigenstrains, it follows that in this example the elastic strain in the region is negative, corresponding to a compression by on the region .

The solutions for the total stress and strain within are given by:

Where is the Eshelby Tensor, whose value for each component is determined only by the geometry of the ellipsoid. The solution demonstrates that the total strain and stress state within the inclusion are uniform. Outside of , the stress decays towards zero with increasing distance away from the inclusion. In the general case, the resulting stresses and strains may be asymmetric, and due to the asymmetry of , the eigenstrain may not be coaxial with the total strain.

Inverse problem

Eigenstrains and the residual stresses that accompany them are difficult to measure (see:Residual stress). Engineers can usually only acquire partial information about the eigenstrain distribution in a material. Methods to fully map out the eigenstrain, called the inverse problem of eigenstrain, are an active area of research. [5] Understanding the total residual stress state, based on knowledge of the eigenstrains, informs the design process in many fields.

Applications

Structural engineering

Residual stresses, e.g. introduced by manufacturing processes or by welding of structural members, reflect the eigenstrain state of the material. [5] This can be unintentional or by design, e.g. shot peening. In either case, the final stress state can affect the fatigue, wear, and corrosion behavior of structural components. [7] Eigenstrain analysis is one way to model these residual stresses.

Composite materials

Since composite materials have large variations in the thermal and mechanical properties of their components, eigenstrains are particularly relevant to their study. Local stresses and strains can cause decohesion between composite phases or cracking in the matrix. These may be driven by changes in temperature, moisture content, piezoelectric effects, or phase transformations. Particular solutions and approximations to the stress fields taking into account the periodic or statistical character of the composite material's eigenstrain have been developed. [2]

Strain engineering

Lattice misfit strains are also a class of eigenstrains, caused by growing a crystal of one lattice parameter on top of a crystal with a different lattice parameter. [8] Controlling these strains can improve the electronic properties of an epitaxially grown semiconductor. [9] See: strain engineering.

See also

Related Research Articles

<span class="mw-page-title-main">Composite material</span> Material made from a combination of two or more unlike substances

A composite material is a material which is produced from two or more constituent materials. These constituent materials have notably dissimilar chemical or physical properties and are merged to create a material with properties unlike the individual elements. Within the finished structure, the individual elements remain separate and distinct, distinguishing composites from mixtures and solid solutions.

In continuum mechanics, the infinitesimal strain theory is a mathematical approach to the description of the deformation of a solid body in which the displacements of the material particles are assumed to be much smaller than any relevant dimension of the body; so that its geometry and the constitutive properties of the material at each point of space can be assumed to be unchanged by the deformation.

<span class="mw-page-title-main">Hooke's law</span> Physical law: force needed to deform a spring scales linearly with distance

In physics, Hooke's law is an empirical law which states that the force needed to extend or compress a spring by some distance scales linearly with respect to that distance—that is, Fs = kx, where k is a constant factor characteristic of the spring, and x is small compared to the total possible deformation of the spring. The law is named after 17th-century British physicist Robert Hooke. He first stated the law in 1676 as a Latin anagram. He published the solution of his anagram in 1678 as: ut tensio, sic vis. Hooke states in the 1678 work that he was aware of the law since 1660.

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.

Linear elasticity is a mathematical model of how solid objects deform and become internally stressed due to prescribed loading conditions. It is a simplification of the more general nonlinear theory of elasticity and a branch of continuum mechanics.

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.

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.

Dynamic modulus is the ratio of stress to strain under vibratory conditions. It is a property of viscoelastic materials.

Elastic energy is the mechanical potential energy stored in the configuration of a material or physical system as it is subjected to elastic deformation by work performed upon it. Elastic energy occurs when objects are impermanently compressed, stretched or generally deformed in any manner. Elasticity theory primarily develops formalisms for the mechanics of solid bodies and materials. The elastic potential energy equation is used in calculations of positions of mechanical equilibrium. The energy is potential as it will be converted into other forms of energy, such as kinetic energy and sound energy, when the object is allowed to return to its original shape (reformation) by its elasticity.

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.

A strain energy density function or stored energy density function is a scalar-valued function that relates the strain energy density of a material to the deformation gradient.

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.

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.

In solid mechanics, the Johnson–Holmquist damage model is used to model the mechanical behavior of damaged brittle materials, such as ceramics, rocks, and concrete, over a range of strain rates. Such materials usually have high compressive strength but low tensile strength and tend to exhibit progressive damage under load due to the growth of microfractures.

The viscous stress tensor is a tensor used in continuum mechanics to model the part of the stress at a point within some material that can be attributed to the strain rate, the rate at which it is deforming around that point.

<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 work.

In the physics of continuous media, spatial dispersion is a phenomenon where material parameters such as permittivity or conductivity have dependence on wavevector. Normally, such a dependence is assumed to be absent for simplicity, however spatial dispersion exists to varying degrees in all materials.

<span class="mw-page-title-main">JCMsuite</span> Simulation software

JCMsuite is a finite element analysis software package for the simulation and analysis of electromagnetic waves, elasticity and heat conduction. It also allows a mutual coupling between its optical, heat conduction and continuum mechanics solvers. The software is mainly applied for the analysis and optimization of nanooptical and microoptical systems. Its applications in research and development projects include dimensional metrology systems, photolithographic systems, photonic crystal fibers, VCSELs, Quantum-Dot emitters, light trapping in solar cells, and plasmonic systems. The design tasks can be embedded into the high-level scripting languages MATLAB and Python, enabling a scripting of design setups in order to define parameter dependent problems or to run parameter scans.

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's a special case of elastic behaviour.

References

  1. Kinoshita, N.; Mura, T. (1971). "Elastic fields of inclusions in anisotropic media". Physica Status Solidi A. 5 (3): 759–768. doi:10.1002/pssa.2210050332.
  2. 1 2 Dvorak, George J. (2013). Micromechanics of Composite Materials. Springer Science. ISBN   978-94-007-4100-3.
  3. 1 2 Mura, Toshio (1987). Micromechanics of Defects in Solids (Second, Revised ed.). Kluwer Academic Publishers. ISBN   978-90-247-3256-2.
  4. Robinson, Kenneth (1951). "Elastic Energy of an Ellipsoidal Inclusion in an Infinite Solid". Journal of Applied Physics. 22 (8): 1045. doi:10.1063/1.1700099.
  5. 1 2 3 4 Jun, Tea-Sung; Korsunsky, Alexander M. (2010). "Evaluation of residual stresses and strains using the Eigenstrain Reconstruction Method". International Journal of Solids and Structures. 47 (13): 1678–1686. doi:10.1016/j.ijsolstr.2010.03.002.
  6. 1 2 Eshelby, John Douglas (1957). "The determination of the elastic field of an ellipsoidal inclusion, and related problems" (PDF). Proceedings of the Royal Society A. 241 (1226): 376–396. doi:10.1098/rspa.1957.0133. S2CID   122550488.
  7. Faghidian, S Ali (2014). "Contents Full Article Content List Abstract IntroductionDetermination of the residual fieldsMathematical theory of reconstructionResults and discussionConclusion References Figures & Tables Article Metrics Related Articles Cite Share Request Permissions Explore More Download PDF Inverse determination of the regularized residual stress and eigenstrain fields due to surface peening". The Journal of Strain Analysis for Engineering Design. 50 (2): 84–91. doi:10.1177/0309324714558326. S2CID   138848957.
  8. Tirry, Wim; Schryvers, Dominique (2009). "Linking a completely three-dimensional nanostrain to a structural transformation eigenstrain". Nature Materials. 8 (9): 752–7. doi:10.1038/nmat2488. PMID   19543276.
  9. Hue, Florian; Hytch, Martin; Bender, Hugo; Houdellier, Florent; Claverie, Alain (2008). "Direct Mapping of Strain in a Strained Silicon Transistor by High-Resolution Electron Microscopy" (PDF). Physical Review Letters. 100 (15): 156602. doi:10.1103/PhysRevLett.100.156602. PMID   18518137. S2CID   42476637.