You are on page 1of 9

Equivalent Plastic Strain The equivalent plastic strain gives a measure of the amount of permanent strain in an engineering body.

The equivalent plastic strain is calculated from the component plastic strain as defined in the Equivalent stress/strain section. Most common engineering materials exhibit a linear stress-strain relationship up to a stress level known as the proportional limit. Beyond this limit, the stress-strain relationship will become nonlinear, but will not necessarily become inelastic. Plastic behavior, characterized by nonrecoverable strain or plastic strain, begins when stresses exceed the material's yield point. Because there is usually little difference between the yield point and the proportional limit, the ANSYS program assumes that these two points are coincident in plasticity analyses.

In order to develop plastic strain, plastic material properties must be defined. You may define plastic material properties by defining either of the following in the Engineering Data:

Bilinear Stress/Strain curve. Mulitlinear Stress/Strain curve.

As has been already mentioned in the previous section, there are two differentapproaches for the synthesis of UFG (Zhu and Liao, 2004), in the present section weattempt to illustrate these approaches in detail. A schematic representation of processingroutes and their effect on final grain size refinement is given in figure 1.1. It is obviousfrom the figure that accumulative roll bonding (ARB) and ECAP, variants of severe plastic deformation, can be effectively used to produce UFG materials. Nanomaterialsfind numerous applications depending upon the structure and property variation obtainedduring processing. A summary of the possible variation in the specific properties of nanomaterials is shown in figure 1.2. This describes the variation when the materialreaches in nanoscale regime. The synthesis route of nanomaterials can be divided intotwo broad approaches which are described in the following sections.

1.1.1

Bottom up approach

In the bottom-up approach of synthesis, atoms, molecules and evennanoparticles can be used as the building blocks for the creation of complex structures.Most of the bottom-up approaches emphasize use of nanopowders for the synthesis of nanostructured materials. For structural applications, the nanopowders need to beconsolidated into bulk nanostructured materials. Examples of these techniques includeinert gas condensation(Gleiter, 1981 and 1989),electrodeposition (Erb et al.,1993), ball milling with subsequent consolidation (Koch and Cho, 1992) and cryomilling with hot isostatic pressing (Luton et al.,1989 and Witkin and Lavernia, 2006).In practice, these

techniques are often limited to the production of fairly small samples that may be usefulfor applications in fields such as electronic devices but are generally not appropriate for largescale structural applications. Furthermore, the finished products from thesetechniques invariably contain some degree of residual porosity and a low level of contamination, which is introduced during the fabrication procedure. Recent research hasshown that large bulk solids, in an essentially fully-dense state, may be produced bycombining cryomilling and hot isostatic pressing with subsequent extrusion (Han et al.,2004) but the operation of this combined procedure is expensive and at present it is noteasily adapted for the production and utilization of structural alloys in large-scaleindustrial applications. 1.1.2 Top-down approach

The top-down processes are effective examples of solid-state processing of materials. In this synthesis approach, coarse-grained materials are refined intonanostructured materials through heavy straining or shock loading. This approachobviates the limitation of small product sizes and also the contamination that is inherentfeatures of materials produced using the bottom-up approach. This has an additionaladvantage that it can be readily applied to a wide range of pre-selected alloys. The firstobservations of the production of UFG microstructures using this approach appeared inthe scientific literature in the early 1990s. (Valiev et al., 1990 and Valiev et al., 1991). Itis important to note that these early publications provided a direct demonstration of theability to employ heavy plastic straining in the production of bulk materials having fairlyhomogeneous and equiaxed microstructures with grain sizes in the submicrometer range. 1.2 Severe plastic deformation Severe plastic deformation term (SPD) is a modified form of intensive plasticdeformation. The term severe plastic deformation was first introduced by Musalimoveand Valiev in 1992, where they described the deformation of an Al-4% Cu-0.5%Zr alloy[R.S. Musalimove and R.Z. Valiev (1992)]. In the last decade, this process established- 2 itself very well as an effective method for the production of bulk ultra fine-grained (UFG)metallic materials (Valiev et al., 1993; Valiev et al., 2000; Valiev, 2004).Severalmethods of SPD are now available for refining the microstructure in order to achievesuperior strength and other properties

Top-down Approach In top-down approach, nano-scale objects are made by processing larger objects in size. Integrated circuit fabrication is an example for top down nanotechnology. Now it has been grown to the level of fabricating nano electromechanical systems (NEMS) where tiny mechanical components such as levers, springs and fluid channels along with electronic circuits are embedded to a tiny chip. The starting materials in these fabrications are relatively large structures such as silicon crystals. Lithography is the technology which has enabled making such tiny chips and there are many types of them such as photo, electron beam and ion beam lithography. In some applications larger scale materials are grinded to the nanometer scale to increase the surface area to volume aspect ratio for more reactivity. Nano gold, nano silver and nano titanium dioxide are such nano materials used in different applications. Carbon nanotube manufacturing process using graphite in an arc oven is another example for top-down approach nanotechnology. Bottom up Approach Bottom-up approach in nanotechnology is making larger nanostructures from smaller building blocks such as atoms and molecules. Self assembly in which desired nano structures are self assembled without any external manipulation. When the object size is getting smaller in nanofabrication, bottom-up approach is an increasingly important complement to top-down techniques. Bottom-up approach nanotechnology can be found from nature, where biological systems have exploited chemical forces to create structures for cells needed for life. Scientists and engineers perform research to imitate this quality of nature to produce small clusters of specific atoms, which can then self assemble into more complex structures. Manufacturing of carbon nanotubes using metal catalyzed polymerization method is a good example for bottom-up approach nanotechnology.

Difference between Top-down and Bottom-up approach in nanotechnology 1. Manufacturing process starts from larger structures in top-down approach where starting building blocks are smaller than the final design in bottom-up approach 2. Bottom-up manufacturing can produce structures with perfect surfaces and edges (not wrinkly and does not contain cavities etc.) though surfaces and edges resulted by top-down manufacturing are not perfect as they are wrinkly or containing cavities. 3. Bottom-up approach manufacturing technologies are newer than top-down manufacturing and expected to be an alternative for it in some applications (example: transistors). 4. Bottom-up approach products have a higher precision accuracy (more control over the material dimensions) and therefore can manufacture smaller structures compared to top-down approach.

5. In top-down approach there is a certain amount of wasted material as some parts are removed from the original structure contrast to bottom-up approach where no material part is removed.

The von Mises yield criterion[1] suggests that the yielding of materials begins when the second deviatoric stress invariant J2 reaches a critical value k. For this reason, it is sometimes called the J2-plasticity or J2 flow theory. It is part of a plasticity theory that applies best to ductile materials, such as metals. Prior to yield, material response is assumed to be elastic. In materials science and engineering the von Mises yield criterion can be also formulated in terms of the von Mises stress or equivalent tensile stress, v, a scalar stress value that can be computed from the stress tensor. In this case, a material is said to start yielding when its von Mises stress reaches a critical value known as the yield strength, y

Deformation in continuum mechanics is the transformation of a body from a reference configuration to a current configuration.[1] A configuration is a set containing the positions of all particles of the body. Contrary to the common definition of deformation, which implies distortion or change in shape, the continuum mechanics definition includes rigid body motions where shape changes do not take place ([1] footnote 4, p. 48). The cause of a deformation is not pertinent to the definition of the term. However, it is usually assumed that a deformation is caused by external loads,[2] body forces (such as gravity or electromagnetic forces), or temperature changes within the body. Strain is a description of deformation in terms of relative displacement of particles in the body. Different equivalent choices may be made for the expression of a strain field depending on whether it is defined in the initial or in the final placement and on whether the metric tensor or its dual is considered. In a continuous body, a deformation field results from a stress field induced by applied forces or is due to changes in the temperature field inside the body. The relation between stresses and induced strains is expressed by constitutive equations, e.g., Hooke's law for linear elastic materials. Deformations which are recovered after the stress field has been removed are called elastic deformations. In this case, the continuum completely recovers its original configuration. On the other hand, irreversible deformations remain even after stresses have been removed. One type of irreversible deformation is plastic deformation, which occurs in material bodies after stresses have attained a certain threshold value known as the elastic limit or yield stress, and are the result of slip, or dislocation mechanisms at the atomic level. Another type of irreversible deformation is viscous deformation, which is the irreversible part of viscoelastic deformation.

In the case of elastic deformations, the response function linking strain to the deforming stress is the compliance tensor of the material. Continuum mechanics

[show]Laws [show]Solid mechanics [show]Fluid mechanics [show]Rheology [show]Scientists vde

[edit] Strain A strain is a normalized measure of deformation representing the displacement between particles in the body relative to a reference length. A general deformation of a body can be expressed in the form where is the reference position of material points in the body. Such a measure does not distinguish between rigid body motions (translations and rotations) and changes in shape (and size) of the body. A deformation has units of length. We could, for example, define strain to be

. Hence strains are dimensionless and are usually expressed as a decimal fraction, a percentage or in parts-per notation. Strains measure how much a given deformation differs locally from a rigid-body deformation.[3]

A strain is in general a tensor quantity. Physical insight into strains can be gained by observing that a given strain can be decomposed into normal and shear components. The amount of stretch or compression along a material line elements or fibers is the normal strain, and the amount of distortion associated with the sliding of plane layers over each other is the shear strain, within a deforming body.[4] This could be applied by elongation, shortening, or volume changes, or angular distortion.[5] The state of strain at a material point of a continuum body is defined as the totality of all the changes in length of material lines or fibers, the normal strain, which pass through that point and also the totality of all the changes in the angle between pairs of lines initially perpendicular to each other, the shear strain, radiating from this point. However, it is sufficient to know the normal and shear components of strain on a set of three mutually perpendicular directions. If there is an increase in length of the material line, the normal strain is called tensile strain, otherwise, if there is reduction or compression in the length of the material line, it is called compressive strain. [edit] Strain measures Depending on the amount of strain, or local deformation, the analysis of deformation is subdivided into three deformation theories:

Finite strain theory, also called large strain theory, large deformation theory, deals with deformations in which both rotations and strains are arbitrarily large. In this case, the undeformed and deformed configurations of the continuum are significantly different and a clear distinction has to be made between them. This is commonly the case with elastomers, plastically-deforming materials and other fluids and biological soft tissue. Infinitesimal strain theory, also called small strain theory, small deformation theory, small displacement theory, or small displacement-gradient theory where strains and rotations are both small. In this case, the undeformed and deformed configurations of the body can be assumed identical. The infinitesimal strain theory is used in the analysis of deformations of materials exhibiting elastic behavior, such as materials found in mechanical and civil engineering applications, e.g. concrete and steel. Large-displacement or large-rotation theory, which assumes small strains but large rotations and displacements.

In each of these theories the strain is then defined differently. The engineering strain is the most common definition applied to materials used in mechanical and structural engineering, which are subjected to very small deformations. On the other hand, for some materials, e.g. elastomers and polymers, subjected to large deformations, the engineering definition of strain is not applicable, e.g. typical engineering strains greater than 1%,[6] thus other more complex definitions of strain are required, such as stretch, logarithmic strain, Green strain, and Almansi strain.

[edit] Engineering strain The Cauchy strain or engineering strain is expressed as the ratio of total deformation to the initial dimension of the material body in which the forces are being applied. The engineering normal strain or engineering extensional strain or nominal strain e of a material line element or fiber axially loaded is expressed as the change in length L per unit of the original length L of the line element or fibers. The normal strain is positive if the material fibers are stretched or negative if they are compressed. Thus, we have

where is the engineering normal strain, L is the original length of the fiber and is the final length of the fiber. The true shear strain is defined as the change in the angle (in radians) between two material line elements initially perpendicular to each other in the undeformed or initial configuration. The engineering shear strain is defined as the tangent of that angle, and is equal to the length of deformation at its maximum divided by the perpendicular length in the plane of force application which sometimes makes it easier to calculate. [edit] Stretch ratio The stretch ratio or extension ratio is a measure of the extensional or normal strain of a differential line element, which can be defined at either the undeformed configuration or the deformed configuration. It is defined as the ratio between the final length and the initial length L of the material line.

The extension ratio is approximately related to the engineering strain by

This equation implies that the normal strain is zero, so that there is no deformation when the stretch is equal to unity. The stretch ratio is used in the analysis of materials that exhibit large deformations, such as elastomers, which can sustain stretch ratios of 3 or 4 before they fail. On the other hand, traditional engineering materials, such as concrete or steel, fail at much lower stretch ratios. [edit] True strain The logarithmic strain , also called natural strain, true strain or Hencky strain. Considering an incremental strain (Ludwik)

the logarithmic strain is obtained by integrating this incremental strain:

where e is the engineering strain. The logarithmic strain provides the correct measure of the final strain when deformation takes place in a series of increments, taking into account the influence of the strain path.[4] [edit] Green strain Main article: Finite strain theory The Green strain is defined as:

[edit] Almansi strain Main article: Finite strain theory The Euler-Almansi strain is defined as

[edit] Normal strain

Two-dimensional geometric deformation of an infinitesimal material element. As with stresses, strains may also be classified as 'normal strain' and 'shear strain' (i.e. acting perpendicular to or along the face of an element respectively). For an isotropic material that obeys Hooke's law, a normal stress will cause a normal strain. Normal strains produce dilations.

Consider a two-dimensional infinitesimal rectangular material element with dimensions , which after deformation, takes the form of a rhombus. From the geometry of the adjacent figure we have Superplasticity the ability of a material to sustain large plastic deformation has been demonstrated in a number of metallic, intermetallic and ceramic systems. Conditions considered necessary for superplasticity are a stable fine-grained microstructure and a temperature higher than 0.5 Tm (where Tm is the melting point of the matrix). Superplastic behaviour is of industrial interest, as it forms the basis of a fabrication method that can be used to produce components having complex shapes from materials that are hard to machine, such as metal matrix composites and intermetallics. Use of superplastic forming may become even more widespread if lower deformation temperatures can be attained.

In order to improve the fracture toughness in ultrafine-grained metals, we investigate the interactions among crack tips, dislocations, and grain boundaries in aluminum bicrystal models containing a crack and 112 tilt grain boundaries using molecular dynamics simulations. The results of previous computer simulations showed that grain refinement makes materials brittle if grain boundaries behave as obstacles to dislocation movement. However, it is actually well known that grain refinement increases fracture toughness of materials. Thus, the role of grain boundaries as dislocation sources should be essential to elucidate fracture phenomena in ultrafine-grained metals. A proposed mechanism to express the improved fracture toughness in ultrafine-grained metals is the disclination shielding effect on the crack tip mechanical field. Disclination shielding can be activated when two conditions are present. First, a transition of dislocation sources from crack tips to grain boundaries must occur. Second, the transformation of grain-boundary structure into a neighboring energetically stable boundary must occur as dislocations are emitted from the grain boundary. The disclination shielding effect becomes more pronounced as antishielding dislocations are continuously emitted from the grain boundary without dislocation emissions from crack tips, and then ultrafine-grained metals can sustain large plastic deformation without fracture with the drastic increase of the mobile dislocation density. Consequently, it can be expected that the disclination shielding effect can improve the fracture toughness in ultrafine-grained metals.

You might also like