Вы находитесь на странице: 1из 23

Viscoelasticity

Viscoelasticity is the property of materials that exhibit both viscous and elastic characteristics when undergoingdeformation. Viscous materials, like honey, resist shear flow and strain linearly with time when a stress is applied. Elastic materials strain instantaneously when stretched and just as quickly return to their original state once the stress is removed. Viscoelastic materials have elements of both of these properties and, as such, exhibit time dependent strain. Whereas elasticity is usually the result of bond stretching along crystallographic planes in an ordered solid, viscosity is the result of the diffusion of atoms or molecules inside an amorphous material.[1]

Background
In the nineteenth century, physicists such as Maxwell, Boltzmann, and Kelvin researched and experimented with creep and recovery of glasses, metals, and rubbers.[2] Viscoelasticity was further examined in the late twentieth century when synthetic polymers were engineered and used in a variety of applications. [2] Viscoelasticity calculations depend heavily on the viscosity variable, . The inverse of is also known as fluidity, . The value of either can be derived as a function of temperature or as a given value (i.e. for a dashpot).[1]

Different types of responses (

) to a change in strain rate (d /dt)

Depending on the change of strain rate versus stress inside a material the viscosity can be categorized as having a linear, non-linear, or plastic response. When a material exhibits a linear response it is categorized as a Newtonian material.[1] In this case the stress is linearly proportional to the strain rate. If the material exhibits a non-linear response to the strain rate, it is categorized as Non-Newtonian fluid. There is also an interesting case where the viscosity decreases as the shear/strain rate remains constant. A material which exhibits this type of behavior is known as thixotropic.[1] In addition, when the stress is independent of this strain rate, the material exhibits plastic deformation.[1] Many viscoelastic materials exhibit rubber like behavior explained by

the thermodynamic theory of polymer elasticity. In reality all materials deviate from Hooke's law in various ways, for example by exhibiting viscous-like as well as elastic characteristics. Viscoelastic materials are those for which the relationship between stress and strain depends on time. Anelastic solids represent a subset of viscoelastic materials: they have a unique equilibrium configuration and ultimately recover fully after removal of a transient load.Some phenomena in viscoelastic materials are:

if the stress is held constant, the strain increases with time (creep); if the strain is held constant, the stress decreases with time (relaxation); the effective stiffness depends on the rate of application of the load; if cyclic loading is applied, hysteresis (a phase lag) occurs, leading to a dissipation of mechanical energy; acoustic waves experience attenuation; rebound of an object following an impact is less than 100%; during rolling, frictional resistance occurs.

a) Applied strain and b) induced stress as functions of time for a viscoelastic material.

All materials exhibit some viscoelastic response. In common metals such as steel or aluminum, as well as in quartz, at room temperature and at small strain, the behavior does not deviate much from linear elasticity. Synthetic polymers, wood, and human tissue as well as metals at high temperature display significant viscoelastic effects. In some applications, even a small viscoelastic response can be significant. To be complete, an analysis or design involving such materials must incorporate their viscoelastic behavior. Knowledge of the viscoelastic response of a material is based on measurement.Some examples of viscoelastic materials include amorphous polymers, semicrystalline polymers, biopolymers, metals at very high temperatures, and bitumen materials. Cracking occurs when the strain is applied quickly and outside of the elastic limit.Ligaments and tendons are viscoelastic, so the extent of the potential damage to them depends both on the velocity of the change of their length as well as on the force applied.

A viscoelastic material has the following properties:

hysteresis is seen in the stress-strain curve. stress relaxation occurs: step constant strain causes decreasing stress creep occurs: step constant stress causes increasing strain

[edit]Elastic

behavior versus viscoelastic behavior

Stress-Strain Curves for a purely elastic material (a) and a viscoelastic material (b). The red area is a hysteresis loop and shows the amount of energy lost (as heat) in a loading and unloading cycle. It is equal to is strain.
[1]

, where

is stress and

Unlike purely elastic substances, a viscoelastic substance has an elastic component and a viscous component. The viscosity of a viscoelastic substance gives the substance a strain rate dependent on time. [1]Purely elastic materials do not dissipate energy (heat) when a load is applied, then removed. [1] However, a viscoelastic substance loses energy when a load is applied, then removed.Hysteresis is observed in the stress-strain curve, with the area of the loop being equal to the energy lost during the loading cycle.[1] Since viscosity is the resistance to thermally activated plastic deformation, a viscous material will lose energy through a loading cycle. Plastic deformation results in lost energy, which is uncharacteristic of a purely elastic material's reaction to a loading cycle.[1] Specifically, viscoelasticity is a molecular rearrangement. When a stress is applied to a viscoelastic material such as a polymer, parts of the long polymer chain change position. This movement or rearrangement is called Creep. Polymers remain a solid material even when these parts of their chains are rearranging in order to accompany the stress, and as this occurs, it creates a back stress in the material. When the back stress is the same magnitude as the applied stress, the material no longer creeps. When the original stress is taken away, the accumulated back stresses will cause the polymer to return to its original form. The material creeps, which gives the prefix visco-, and the material fully recovers, which gives the suffix -elasticity.[2]

Types of viscoelasticity
Linear viscoelasticity is when the function is separable in both creep response and load. All linear viscoelastic models can be represented by a Volterra equation connecting stress and strain:

or

where

t is time
is stress is strain and are instantaneous elastic moduli for creep and relaxation

K(t) is the creep function F(t) is the relaxation function


Linear viscoelasticity is usually applicable only for small deformations. Nonlinear viscoelasticity is when the function is not separable. It usually happens when thedeformations are large or if the material changes its properties under deformations. An anelastic material is a special case of a viscoelastic material: an anelastic material will fully recover to its original state on the removal of load.

Dynamic modulus
Main article: Dynamic modulus Viscoelasticity is studied using dynamic mechanical analysis, applying a small oscillatory stress and measuring the resulting strain.

Purely elastic materials have stress and strain in phase, so that the response of one caused by
the other is immediate.

In purely viscous materials, strain lags stress by a 90 degree phase lag.

Viscoelastic materials exhibit behavior somewhere in the middle of these two types of material,
exhibiting some lag in strain. Complex Dynamic modulus G can be used to represent the relations between the oscillating stress and strain:

where

is the storage modulus and

is the loss modulus:

where them.

and

are the amplitudes of stress and strain and

is the phase shift between

Constitutive models of linear viscoelasticity


Viscoelastic materials, such as amorphous polymers, semicrystalline polymers, and biopolymers, can be modeled in order to determine their stress or strain interactions as well as their temporal dependencies. These models, which include the Maxwell model, the Kelvin-Voigt model, and theStandard Linear Solid Model, are used to predict a material's response under different loading conditions. Viscoelastic behavior has elastic and viscous components modeled as linear combinations of springs and dashpots, respectively. Each model differs in the arrangement of these elements, and all of these viscoelastic models can be equivalently modeled as electrical circuits. In an equivalent electrical circuit, stress is represented by voltage, and the derivative of strain (velocity) by current. The elastic modulus of a spring is analogous to a circuit's capacitance (it stores energy) and the viscosity of a dashpot to a circuit's resistance (it dissipates energy). The elastic components, as previously mentioned, can be modeled as springs of elastic constant E, given the formula:

where is the stress, E is the elastic modulus of the material, and is the strain that occurs under the given stress, similar to Hooke's Law. The viscous components can be modeled as dashpots such that the stress-strain rate relationship can be given as,

where is the stress, is the viscosity of the material, and d/dt is the time derivative of strain. The relationship between stress and strain can be simplified for specific stress rates. For high stress states/short time periods, the time derivative components of the stress-strain relationship dominate. A dashpot resists changes in length, and in a high stress state it can be approximated as a rigid rod. Since a rigid rod cannot be stretched past its original length, no strain is added to the system[3] Conversely, for low stress states/longer time periods, the time derivative components are negligible and the dashpot can be effectively removed from the system - an "open" circuit. As a result, only the spring connected in parallel to the dashpot will contribute to the total strain in the system[3]

Maxwell model
Main article: Maxwell material

Maxwell model

The Maxwell model can be represented by a purely viscous damper and a purely elastic spring connected in series, as shown in the diagram. The model can be represented by the following equation:

. Under this model, if the material is put under a constant strain, the stresses gradually relax, When a material is put under a constant stress, the strain has two components. First, an elastic component occurs instantaneously, corresponding to the spring, and relaxes immediately upon release of the stress. The second is a viscous component that grows with time as long as the stress is applied. The Maxwell model predicts that stress decays exponentially with time, which is accurate for most polymers. One limitation of this model is that it does not predict creep accurately. The Maxwell model for creep or constant-stress conditions postulates that strain will increase linearly with time. However, polymers for the most part show the strain rate to be decreasing with time.[2] Application to soft solids:thermoplastic polymers in the vicinity of their melting temperature, fresh concrete (neglecting its aging), numerous metals at a temperature close to their melting point.

Schematic representation of KelvinVoigt model

The KelvinVoigt model, also known as the Voigt model, consists of a Newtonian damper and Hookean elastic spring connected in parallel, as shown in the picture. It is used to explain the creep behaviour of polymers. The constitutive relation is expressed as a linear first-order differential equation:

This model represents a solid undergoing reversible, viscoelastic strain. Upon application of a constant stress, the material deforms at a decreasing rate, asymptotically approaching the steadystate strain. When the stress is released, the material gradually relaxes to its undeformed state. At constant stress (creep), the Model is quite realistic as it predicts strain to tend to /E as time continues to infinity. Similar to the Maxwell model, the KelvinVoigt model also has limitations. The model is extremely good with modelling creep in materials, but with regards to relaxation the model is much less accurate. Applications: organic polymers, rubber, wood when the load is not too high.

Standard linear solid model


Main article: Standard linear solid model

Schematic representation of the Standard Linear Solid model.

The Standard Linear Solid Model effectively combines the Maxwell Model and a Hookean spring in parallel. A viscous material is modeled as a spring and a dashpot in series with each other, both of which are in parallel with a lone spring. For this model, the governing constitutive relation is:

Under a constant stress, the modeled material will instantaneously deform to some strain, which is the elastic portion of the strain, and after that it will continue to deform and asymptotically approach a steady-state strain. This last portion is the viscous part of the strain. Although the Standard Linear Solid Model is more accurate than the Maxwell and Kelvin-Voigt models in predicting material responses, mathematically it returns inaccurate results for strain under specific loading conditions and is rather difficult to calculate.

[edit]Generalized

Maxwell Model

Main article: Generalized Maxwell Model

Schematic of Maxwell-Wiechert Model

The Generalized

Maxwell

model also
[4][5]

known

as

the MaxwellWiechert

model(after James Clerk Maxwell and E Wiechert

) is the most general form of the linear model for

viscoelasticity. It takes into account that the relaxation does not occur at a single time, but at a distribution of times. Due to molecular segments of different lengths with shorter ones contributing less than longer ones, there is a varying time distribution. The Wiechert model shows this by having as many spring dashpot Maxwell elements as are necessary to accurately represent the distribution. The figure on the right shows the generalised Wiechert model
[6]

Applications : metals and alloys at temperatures lower than one

quarter of their absolute melting temperature (expressed in K).

Prony series
In a one-dimensional relaxation test, the material is subjected to a sudden strain that is kept constant over the duration of the test, and the stress is measured over time. The initial stress is due to the elastic response of the material. Then, the stress relaxes over time due to the viscous effects in the material. Typically, either a tensile, compressive, bulk compression, or shear strain is applied. The resulting stress vs. time data can be fitted with a number of equations, called models. Only the notation changes

depending of the type of strain applied: tensile-compressive relaxation is denoted bulk is denoted . The Prony series for the shear relaxation is

, shear is denoted

where

is the long term modulus once the material is totally relaxed,

are the

relaxation times (not to be confused with

in the diagram); the higher their values, the longer it takes for

the stress to relax. The data is fitted with the equation by using a minimization algorithm that adjust the parameters ( ) to minimize the error between the predicted and data values .[7] An alternative form is obtained noting that the elastic modulus is related to the long term modulus by

Therefore,

This form is convenient when the elastic shear modulus

is obtained from data

independent from the relaxation data, and/or for computer implementation, when it is desired to specify the elastic properties separately from the viscous properties, as in .[8] A creep experiment is usually easier to perform than a relaxation one, so most data is available as (creep) compliance vs. time. [9] Unfortunately, there is no known closed form for the (creep) compliance in terms of the coefficient of the Prony series. So, if one has creep data, it is not easy to get the coefficients of the (relaxation) Prony series, which are needed for example in.[8] An expedient way to obtain these coefficients is the following. First, fit the creep data with a model that has closed form solutions in both compliance and relaxation; for example the Maxwell-Kelvin model (eq. 7.18-7.19) in[10] or the Standard Solid Model (eq. 7.20-7.21) in [10] (section 7.1.3). Once the parameters of the creep model are known, produce relaxation pseudo-data with the conjugate relaxation model for the same times of the original data. Finally, fit the pseudo data with the Prony series.

Effect of temperature on viscoelastic behavior


The secondary bonds of a polymer constantly break and reform due to thermal motion. Application of a stress favors some conformations over others, so the molecules of the polymer will gradually "flow" into the favored conformations over time.[11] Because thermal motion is one factor contributing to the deformation of polymers, viscoelastic properties change with increasing or decreasing temperature. In most cases, the creep modulus, defined as the ratio of applied stress to the time-dependent strain, decreases with increasing temperature. Generally speaking, an increase in temperature correlates to a logarithmic decrease in the time required to impart equal strain under a constant stress. In other words, it takes less

work to stretch a viscoelastic material an equal distance at a higher temperature than it does at a lower temperature.

Viscoelastic creep
Main article: Creep (deformation)

a) Applied stress and b) induced strain (b) as functions of time over a short period for a viscoelastic material.

When subjected to a step constant stress, viscoelastic materials experience a time-dependent increase in strain. This phenomenon is known as viscoelastic creep. At a time , a viscoelastic material is loaded with a constant stress that is maintained for a sufficiently long

time period. The material responds to the stress with a strain that increases until the material ultimately fails, if it is a viscoelastic liquid. If, on the other hand, it is a viscoelastic solid, it may or may not fail depending on the applied stress versus the material's ultimate resistance. When the stress is maintained for a shorter time period, the material undergoes an initial strain until a time immediately decreases (discontinuity) then gradually decreases at times , after which the strain to a residual strain.

Viscoelastic creep data can be presented by plotting the creep modulus (constant applied stress divided by total strain at a particular time) as a function of time. [12] Below its critical stress, the viscoelastic creep modulus is independent of stress applied. A family of curves describing strain versus time response to various applied stress may be represented by a single viscoelastic creep modulus versus time curve if the applied stresses are below the material's critical stress value. Viscoelastic creep is important when considering long-term structural design. Given loading and temperature conditions, designers can choose materials that best suit component lifetimes.

Measuring viscoelasticity

Though there are many instruments that test the mechanical and viscoelastic response of materials,broadband viscoelastic spectroscopy (BVS) and resonant ultrasound spectroscopy (RUS) are more commonly used to test viscoelastic behavior because they can be used above and below ambient temperatures and are more specific to testing viscoelasticity. These two instruments employ a damping mechanism at various frequencies and time ranges with no appeal to timetemperature superposition.
[13]

Using BVS and RUS to study the mechanical properties of materials is important to understanding how

a material exhibiting viscoelasticity will perform.[13]

Hysteresis is the dependence of a system not only on its current environment but also on its past environment. This dependence arises because the system can be in more than one internal state. To predict its future development, either its internal state or its history must be known.[1] If a given input alternately increases and decreases, the output tends to form a loop as in the figure. However, loops may also occur because of a dynamic lag between input and output. Often, this effect is also referred to as hysteresis, orrate-dependent hysteresis. This effect disappears as the input changes more slowly, so many experts do not regard it as true hysteresis. Hysteresis occurs in ferromagnetic materials andferroelectric materials, as well as in the deformation of some materials (such as rubber bands and shape-memory alloys) in response to a varying force. In natural systems hysteresis is often associated withirreversible thermodynamic change. Many artificial systems are designed to have hysteresis: for example, in thermostats and Schmitt triggers, hysteresis is produced by positive feedback to avoid unwanted rapid switching. Hysteresis has been identified in many other fields, including economicsand biology.

Hysteresis in mechanics
Elastic hysteresis of an idealized rubber band.The area in the centre of the hysteresis loop is the energy dissipated due to material plasticity

Elastic hysteresis
In the elastic hysteresis of rubber, the area in the centre of a hysteresis loop is the energy dissipated due to material plasticity. Elastic hysteresis was one of the first types of hysteresis to be examined.[8][9] A simple way to understand it is in terms of a rubber band with weights attached to it. If the top of a rubber band is hung on a hook and small weights are attached to the bottom of the band one at a time, it will get longer. As more weights are loaded onto it, the band will continue to

extend because the force the weights are exerting on the band is increasing. When each weight is taken off, or unloaded, it will get shorter as the force is reduced. As the weights are taken off, each weight that produced a specific length as it was loaded onto the band now produces a slightly longer length as it is unloaded. This is because the band does not obey Hooke's law perfectly. The hysteresis loop of an idealized rubber band is shown in the figure. In one sense the rubber band was harder to stretch when it was being loaded than when it was being unloaded. In another sense, as one unloads the band, the cause (the force of the weights) lags behind the effect (the length) because a smaller value of weight produces the same length. In another sense more energy was required during the loading than the unloading; that energy must have gone somewhere, it was dissipated or "lost" as heat. Elastic hysteresis is more pronounced when the loading and unloading is done quickly than when it is done slowly.[10] Some materials such as hard metals don't show elastic hysteresis under a moderate load, whereas other hard materials like granite and marble do. Materials such as rubber exhibit a high degree of elastic hysteresis. A word of caution: rubber behaves like a gas. When the rubber band is stretched it heats up. If it is suddenly released, the rubber cools down, very easy to perceive just by touching. So, there is a large hysteresis from the thermal exchange with the environment and a smaller hysteresis due to internal friction within the rubber. This proper, intrinsic hysteresis could be measured only if adiabatic isolation of the rubber band is imposed. Small vehicle suspensions using rubber (or other elastomers) can achieve the dual function of springing and damping because rubber, unlike metal springs, has pronounced hysteresis and does not return all the absorbed compression energy on the rebound. Mountain bikes have frequently made use of elastomer suspension, as did the original Mini car.

Viscoelastic properties of materials


Dynamic mechanical analysis (abbreviated DMA, also known as dynamic mechanical spectroscopy) is a technique used to study and characterize materials. It is most useful for studying the viscoelastic behavior ofpolymers. A sinusoidal stress is applied and the strain in the material is measured, allowing one to determine thecomplex 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.

Polymers composed of long molecular chains have unique viscoelastic properties, which combine the characteristics of elastic solids andNewtonian fluids. The classical theory of elasticity describes the mechanical properties of elastic solids where stress is proportional to strain in small deformations. Such response of stress is independent of strain rate. The classical theory of hydrodynamics describes the properties of viscous fluid, for which the response of stress is dependent on strain rate.[1] This solidlike and liquidlike behavior of polymer can be modeled mechanically with combinations of springs and dashpots.[2] [edit]Dynamic

moduli of polymers

The viscoelastic property of a polymer is studied by dynamic mechanical analysis where a sinusoidal force (stress ) is applied to a material and the resulting displacement (strain) is measured. For a perfectly elastic solid, the resulting strain and the stress will be perfectly in phase. For a purely viscous fluid, there will be a 90 degree phase lag of strain with respect to stress.[3] Viscoelastic polymers have the characteristics in between where some phase lag will occur during DMA tests.[3] where is frequency of strain oscillation, is time, is phase lag between stress and strain. The storage modulus measures the stored energy, representing the elastic portion, and the loss modulus measures the energy dissipated as heat, representing the viscous portion.[3] The tensile storage and loss moduli are defined as follows: Stress: Strain:
[3]

Storage Modulus: Loss Modulus: Phase angle:, Tan (delta): and and . as follows:

Similarly we also define shear storage and loss moduli, Complex variables can be used to express the moduli

where

Stress relaxation and creep are the same phenomenon except that one looks at it from a different perspective. Polymers under constant stress creep because the material is constantly attempting to relax the stress. In the same vein, plastics stress-relax under constant strain because the stretched polymer chains are constantly attempting to creep or flow to relieve the stress caused by the stretching. The ideal way to generate stress relaxation data is to take a specimen, put it under a particular strain in a tensile machine and monitor the stress as it decays with time. The problem is that the machine is tied up for long periods of time --- but, it is by far the best way to develop accurate stress relaxation data.

An alternative approach is to use the creep modulus data to estimate the stress relaxation at some point in time. A word of caution, however: compare the apparent creep modulus to the original short term modulus of the material --- dont compare the time-dependent creep modulus to the modulus at 10 hours, the typical origin for a logarithmic curve (this is not the modulus at zero time, but the creep modulus at 1 hour --- by this time most polymers will have relaxed about 20% of the initial load). Keep in mind that flexural creep data may be somewhat conservative, and somewhat misleading at strains significantly beyond 5%.
0

In a paper entitled, Pull-Out Forces on Joints in Polyethylene Pipe Systems, the authors, J.L. Husted and D.M. Thompson, listed the following relationship between apparent as (Instantaneous) modulus % and time: of Modulus

Modulus Time Initial

6 1 10 4 6 1.1 11 50

min hour hours days weeks years years years 1.0 10 100 1000 10,000 100,000 438,000

0.1

hour hour hours hours hours hours hours hours

100% 80% 67% 51% 39% 30% 24% 21%

These data were developed for polyethylene, however, interestingly enough, these same estimated relaxations have worked well for predicting stress relaxation in other plastic materials including other polyolefins, acetals, and PVC. Even more useful is the fact that these same apparent relaxations work well as a guideline for estimating the relaxation of polymers at various temperatures, when given the short-term modulus of the material at that temperature.

The shape and magnitude of the stress-strain curve of a metal will depend on its composition, heat treatment, prior history of plastic deformation, and the strain rate, temperature, and state of stress imposed during the testing. The parameters, which are used to describe the stress-strain curve of a metal, are the tensile strength, yield strength or yield point, percent elongation, and reduction of area. The first two are strength parameters; the last two indicate ductility. An example of the engineering stress-strain curve for a typical engineering alloy is shown in Figure 1. From it some very important properties can be determined. The elastic modulus, the yield strength, the ultimate tensile strength, and the fracture strain are all clearly exhibited in an accurately constructed stress strain curve.

Figure 1: An example of the engineering stress strain curve for a typical engineering alloy The elastic modulus, E (Youngs modulus) is the slope of the elastic portion of the curve (the steep, linear region) because E is the proportionality constant relating stress and strain during elastic deformation: = E. The 0.2% offset yield strength is the stress value, 0.2%YS of the intersection of a line (called the offset) constructed parallel to the elastic portion of the curve but offset to the right by a strain of 0.002. It represents the onset of plastic deformation.

The ultimate tensile strength is the engineering stress value or

ut

, at the maximum of the engineering stress-strain curve. It

represents the maximum load, for that original area, that the sample can sustain without undergoing the instability of necking, which will lead inexorably to fracture. The fracture strain is the engineering strain value at which fracture occurred. At the outset, though, a clear distinction must be made between a true stress-true strain curve and an engineering stressengineering strain curve. The difference is shown in Figure 2, which are plotted, on the same axes, the stress-strain curve and engineering stress-strain curve for the same material. The difference is also evident in the definitions of true stress-true strain and engineering stress-engineering strain.

Figure 2: Comparison of engineering and true stress-strain curves The engineering stress is the load borne by the sample divided by a constant, the original area. The true stress is the load borne by the

sample divided by a variable the instantaneous area. Note that the true stress always rises in the plastic, whereas the engineering stress rises and then falls after going through a maximum. The maximum represents a significant difference between the engineering stress-strain curve and the true stress-strain curve. In the engineering stress-strain curve, this point indicates the beginning of necking. The ultimate tensile strength is the maximum load measured in the tension test divided by the original area.

Deformation (engineering)

Compressive stress results in deformation which shortens the object but also expands it outwards.

In materials science, deformation is a change in the shape or size of an object due to an applied force (the deformation energy in this case is transferred through work) or a change in temperature (the deformation energy in this case is transferred through heat). The first case can be a result of tensile (pulling) forces,compressive (pushing) forces, shear, bending or torsion (twisting). In the second case, the most significant factor, which is determined by the temperature, is the mobility of the structural defects such as grain boundaries, point vacancies, line and screw dislocations, stacking faults and twins in both crystalline and non-crystalline solids. The movement or displacement of such mobile defects is thermally activated, and thus limited by the rate of atomic diffusion. Deformation is often described as strain.[1][2] As deformation occurs, internal inter-molecular forces arise that oppose the applied force. If the applied force is not too large these forces may be sufficient to completely resist the applied force, allowing the object to assume a new equilibrium state and to return to its original state when the load is removed. A larger applied force may lead to a permanent deformation of the object or even to itsstructural failure. In the figure it can be seen that the compressive loading (indicated by the arrow) has caused deformation in the cylinder so that the original shape (dashed lines) has changed (deformed) into one with bulging sides. The sides bulge because the material, although strong enough to not crack or otherwise fail, is not strong enough to support the load without change, thus the material is forced out laterally. Internal forces (in this case at right angles to the deformation) resist the applied load. The concept of a rigid body can be applied if the deformation is negligible.

Types of deformation
Depending on the type of material, size and geometry of the object, and the forces applied, various types of deformation may result. The image to the right shows the engineering stress vs. strain diagram for a typical ductile material such as steel. Different deformation modes may occur under different conditions, as can be depicted using a deformation mechanism map.

Typical stress vs. strain diagram with the various stages of deformation.

Elastic deformation
This type of deformation is reversible. Once the forces are no longer applied, the object returns to its original shape.Elastomers andshape memorymetals such asNitinol exhibit large elastic deformation ranges, as doesrubber. However elasticity is nonlinear in these materials. Normal metals, ceramics and most crystals show linear elasticity and a smaller elastic range. Linear elastic deformation is governed by Hooke's law, which states:

Where is the applied stress, is a material constant called Young's modulus, and is the resulting strain. This relationship only applies in the elastic range and indicates that the slope of the stress vs. strain curve can be used to find Young's modulus. Engineers often use this calculation in tensile tests. The elastic range ends when the material reaches its yield strength. At this point plastic deformation begins. Note that not all elastic materials undergo linear elastic deformation; some, such as concrete, gray cast iron, and many polymers, respond nonlinearly. For these materials Hooke's law is inapplicable.[3]

Plastic deformation
See also: Plasticity (physics) This type of deformation is irreversible. However, an object in the plastic deformation range will first have undergone elastic deformation, which is reversible, so the object will return part way to its original shape. Soft thermoplastics have a rather large plastic deformation range as do ductile metals such ascopper, silver, and gold. Steel does, too, but not cast iron. Hard thermosetting plastics, rubber, crystals, and ceramics have minimal plastic deformation ranges. One material with a large plastic deformation range is wet chewing gum, which can be stretched dozens of times its original length.

Under tensile stress plastic deformation is characterized by a strain hardening region and a neckingregion and finally, fracture (also called rupture). During strain hardening the material becomes stronger through the movement of atomic dislocations. The necking phase is indicated by a reduction in cross-sectional area of the specimen. Necking begins after the ultimate strength is reached. During necking, the material can no longer withstand the maximum stress and the strain in the specimen rapidly increases. Plastic deformation ends with the fracture of the material.

Metal fatigue
Another deformation mechanism is metal fatigue, which occurs primarily in ductile metals. It was originally thought that a material deformed only within the elastic range returned completely to its original state once the forces were removed. However, faults are introduced at the molecular level with each deformation. After many deformations, cracks will begin to appear, followed soon after by a fracture, with no apparent plastic deformation in between. Depending on the material, shape, and how close to the elastic limit it is deformed, failure may require thousands, millions, billions, or trillions of deformations. Metal fatigue has been a major cause of aircraft failure, such as the De Havilland Comet, especially before the process was well understood. There are two ways to determine when a part is in danger of metal fatigue; either predict when failure will occur due to the material/force/shape/iteration combination, and replace the vulnerable materials before this occurs, or perform inspections to detect the microscopic cracks and perform replacement once they occur. Selection of materials not likely to suffer from metal fatigue during the life of the product is the best solution, but not always possible. Avoiding shapes with sharp corners limits metal fatigue by reducing stress concentrations, but does not eliminate it.

Compressive failure
Usually, compressive stress applied to bars, columns, etc. leads to shortening. Loading a structural element or a specimen will increase the compressive stress until the reach ofcompressive strength. According to the properties of the material, failure will occur as yield for materials with ductile behavior (most metals, some soils and plastics) or as rupture for brittle behavior (geomaterials, cast iron, glass, etc.). In long, slender structural elements such as columns or truss bars an increase of compressive force F leads to structural failure due to buckling at lower stress than the compressive strength.

Fracture

Diagram of a stress-strain curve, showing the relationship between stress (force applied) and strain (deformation) of a ductile metal.

This type of deformation is also irreversible. A break occurs after the material has reached the end of the elastic, and then plastic, deformation ranges. At this point forces accumulate until they are sufficient to cause a fracture. All materials will eventually fracture, if sufficient forces are applied.

Misconceptions
A popular misconception is that all materials that bend are "weak" and those that don't are "strong." In reality, many materials that undergo large elastic and plastic deformations, such as steel, are able to absorb stresses that would cause brittle materials, such as glass, with minimal plastic deformation ranges, to break.[4]

Fracture strength

Stress vs. strain curve typical of aluminum 1. Ultimate tensile strength 2. Yield strength 3. Proportional limit stress 4. Fracture 5. Offset strain (typically 0.2%)

Fracture strength, also known as breaking strength, is the stress at which a specimen fails via fracture. [1] This is usually determined for a given specimen by a tensile test, which charts the stress-strain curve (see image). The final recorded point is the fracture strength. Ductile materials have a fracture strength lower than theultimate tensile strength (UTS), whereas in brittle materials the fracture strength is equivalent to the UTS.[1] If a ductile material reaches its ultimate tensile strength in a load-controlled situation,[Note 1] it will continue to deform, with no additional load application, until it ruptures. However, if the loading is displacement-controlled,[Note 2] the deformation of the material may relieve the load, preventing rupture. If the stress-strain curve is plotted in terms of true stressand true strain the curve will always slope upwards and never reverse, as true stress is corrected for the decrease in cross-sectional area. The true stress on the

material at the time of rupture is known as the breaking strength. This is the maximum stress on the true stress-strain curve, given by point 3 on curve B.

Types
Brittle fracture

Brittle fracture in glass.

Fracture of an Aluminum Crank Arm. Bright: Brittle fracture. Dark: Fatigue fracture.

In brittle fracture, no apparent plastic deformationtakes place before fracture. In brittle crystalline materials, fracture can occur by cleavage as the result of tensile stress acting normal to crystallographic planes with low bonding (cleavage planes). Inamorphous solids, by contrast, the lack of a crystalline structure results in a conchoidal fracture, with cracks proceeding normal to the applied tension. The theoretical strength of a crystalline material is (roughly)

where: is the Young's modulus of the material, is the surface energy, and

is the equilibrium distance between atomic centers. On the other hand, a crack introduces a stress concentration modeled by

(For sharp cracks) where: is the loading stress, is half the length of the crack, and is the radius of curvature at the crack tip. Putting these two equations together, we get

Looking closely, we can see that sharp cracks (small strength of the material.

) and large defects (large

) both lower the fracture

Recently, scientists have discovered supersonic fracture, the phenomenon of crack motion faster than the speed of sound in a material [2]. This phenomenon was recently also verified by experiment of fracture in rubber-like materials.

Ductile fracture

Ductile failure of a specimen strained axially.

Schematic representation of the steps in ductile fracture (in pure tension).

In ductilefracture, extensive plastic deformation (necking) takes place before fracture. The termsrupture or ductile rupture describe the ultimate failure oftough ductile materials loaded in tension. Rather than cracking, the material "pulls apart," generally leaving a rough surface. In this case there is slow propagation and an absorption of a large amount energy before fracture.[citation needed] Many ductile metals, especially materials with high purity, can sustain very large deformation of 50100% or more strain before fracture under favorable loading condition and environmental condition. The strain at which the fracture happens is controlled by the purity of the materials. At room temperature, pure iron can undergo deformation up to 100% strain before breaking, while cast iron or high-carbon steels can barely sustain 3% of strain.[citation needed] Because ductile rupture involves a high degree of plastic deformation, the fracture behavior of a propagating crack as modeled above changes fundamentally. Some of the energy from stress concentrations at the crack tips is dissipated by plastic deformation before the crack actually propagates. The basic steps are: void formation, void coalescence (also known as crack formation), crack propagation, and failure, often resulting in a cup-and-cone shaped failure surface.

Crack separation modes

The three fracture modes.

There are three ways of applying a force to enable a crack to propagate: Mode I crack Opening mode (a tensile stress normal to the plane of the crack) Mode II crack Sliding mode (a shear stress acting parallel to the plane of the crack and perpendicular to the crack front) Mode III crack Tearing mode (a shear stress acting parallel to the plane of the crack and parallel to the crack front)

For more information, see fracture mechanics. Crack initiation and propagation accompany fracture. The manner through which the crack propagates through the material gives great insight into the mode of fracture. In ductile materials (ductile fracture), the crack moves slowly and is accompanied by a large amount of plastic deformation. The crack will usually not extend unless an increased stress is applied. On the other hand, in dealing with brittle fracture, cracks spread very rapidly with little or no plastic deformation. The cracks that propagate in a brittle material will continue to

grow and increase in magnitude once they are initiated. Another important mannerism of crack propagation is the way in which the advancing crack travels through the material. A crack that passes through the grains within the material is undergoing transgranular fracture. However, a crack that propagates along the grain boundaries is termed an intergranular fracture.

Structural failure refers to loss of the load-carrying capacity of a component or member within a structure or of the structure itself. Structural failure is initiated when the materialis stressed to its strength limit, thus causing fracture or excessive deformations. In a well-designed system, a localized failure should not cause immediate or even progressive collapse of the entire structure. Ultimate failure strength is one of thelimit states that must be accounted for in structural engineering and structural design.

Вам также может понравиться