Standard linear solid model

Last updated

The standard linear solid (SLS), also known as the Zener model after Clarence Zener, [1] is a method of modeling the behavior of a viscoelastic material using a linear combination of springs and dashpots to represent elastic and viscous components, respectively. Often, the simpler Maxwell model and the Kelvin–Voigt model are used. These models often prove insufficient, however; the Maxwell model does not describe creep or recovery, and the Kelvin–Voigt model does not describe stress relaxation. SLS is the simplest model that predicts both phenomena.

Contents

Definition

Materials undergoing strain are often modeled with mechanical components, such as springs (restorative force component) and dashpots (damping component).

Connecting a spring and damper in series yields a model of a Maxwell material while connecting a spring and damper in parallel yields a model of a Kelvin–Voigt material. [2] In contrast to the Maxwell and Kelvin–Voigt models, the SLS is slightly more complex, involving elements both in series and in parallel. Springs, which represent the elastic component of a viscoelastic material, obey Hooke's law:

where σ is the applied stress, E is the Young's modulus of the material, and ε is the strain. The spring represents the elastic component of the model's response. [2]

Dashpots represent the viscous component of a viscoelastic material. In these elements, the applied stress varies with the time rate of change of the strain:

where η is viscosity of the dashpot component.

Solving the model

In order to model this system, the following physical relations must be realized:

For parallel components: , and . [2]

For series components: , and . [2]

Maxwell representation

Standard Linear Solid model, Maxwell representation SLS.svg
Standard Linear Solid model, Maxwell representation

This model consists of two systems in parallel. The first, referred to as the Maxwell arm, contains a spring () and dashpot (viscosity ) in series. [2] The other system contains only a spring ().

These relationships help relate the various stresses and strains in the overall system and the Maxwell arm:

where the subscripts , , and refer to Maxwell, dashpot, spring one and spring two, respectively.

Using these relationships, their time derivatives, and the above stress-strain relationships for the spring and dashpot elements, the system can be modeled as follows:

[3]

The equation can also be expressed as:

or, in dot notation:

The relaxation time, , is different for each material and is equal to

Kelvin-Voigt representation

Standard Linear Solid model, Kelvin representation SLS2.svg
Standard Linear Solid model, Kelvin representation

This model consists of two systems in series. The first, referred to as the Kelvin arm, contains a spring () and dashpot (viscosity ) in parallel. The other system contains only a spring ().

These relationships help relate the various stresses and strains in the overall system and the Kelvin arm:

where the subscripts , , ,and refer to Kelvin, dashpot, spring one, and spring two, respectively.

Using these relationships, their time derivatives, and the above stress-strain relationships for the spring and dashpot elements, the system can be modeled as follows:

or, in dot notation:

The retardation time, , is different for each material and is equal to

Model characteristics

Comparison of creep and stress relaxation for three and four element models Comparison three four element models.svg
Comparison of creep and stress relaxation for three and four element models

The standard linear solid model combines aspects of the Maxwell and Kelvin–Voigt models to accurately describe the overall behavior of a system under a given set of loading conditions. The behavior of a material applied to an instantaneous stress is shown as having an instantaneous component of the response. Instantaneous release of a stress also results in a discontinuous decrease in strain, as is expected. The shape of the time-dependent strain curve is true to the type of equation that characterizes the behavior of the model over time, depending upon how the model is loaded.

Although this model can be used to accurately predict the general shape of the strain curve, as well as behavior for long time and instantaneous loads, the model lacks the ability to accurately model material systems numerically.

The fluid model equivalent to the standard linear solid model includes a dashpot in series with the Kelvin–Voigt model and is called the Jeffreys model. [4]

See also

Related Research Articles

<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.

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.

Hemorheology, also spelled haemorheology, or blood rheology, is the study of flow properties of blood and its elements of plasma and cells. Proper tissue perfusion can occur only when blood's rheological properties are within certain levels. Alterations of these properties play significant roles in disease processes. Blood viscosity is determined by plasma viscosity, hematocrit and mechanical properties of red blood cells. Red blood cells have unique mechanical behavior, which can be discussed under the terms erythrocyte deformability and erythrocyte aggregation. Because of that, blood behaves as a non-Newtonian fluid. As such, the viscosity of blood varies with shear rate. Blood becomes less viscous at high shear rates like those experienced with increased flow such as during exercise or in peak-systole. Therefore, blood is a shear-thinning fluid. Contrarily, blood viscosity increases when shear rate goes down with increased vessel diameters or with low flow, such as downstream from an obstruction or in diastole. Blood viscosity also increases with increases in red cell aggregability.

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.

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. The model was developed independently by the British physicist Lord Kelvin in 1865 and by the German physicist Woldemar Voigt in 1890.

A Burgers material is a viscoelastic material having the properties both of elasticity and viscosity. It is named after the Dutch physicist Johannes Martinus Burgers.

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

The upper-convected Maxwell (UCM) model is a generalisation of the Maxwell material for the case of large deformations using the upper-convected time derivative. The model was proposed by James G. Oldroyd. The concept is named after James Clerk Maxwell.

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.

<span class="mw-page-title-main">Covariant formulation of classical electromagnetism</span> Ways of writing certain laws of physics

The covariant formulation of classical electromagnetism refers to ways of writing the laws of classical electromagnetism in a form that is manifestly invariant under Lorentz transformations, in the formalism of special relativity using rectilinear inertial coordinate systems. These expressions both make it simple to prove that the laws of classical electromagnetism take the same form in any inertial coordinate system, and also provide a way to translate the fields and forces from one frame to another. However, this is not as general as Maxwell's equations in curved spacetime or non-rectilinear coordinate systems.

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

In materials science, stress relaxation is the observed decrease in stress in response to strain generated in the structure. This is primarily due to keeping the structure in a strained condition for some finite interval of time hence causing some amount of plastic strain. This should not be confused with creep, which is a constant state of stress with an increasing amount of strain.

<span class="mw-page-title-main">Generalized Maxwell model</span>

The generalized Maxwell model also known as the Maxwell–Wiechert model is the most general form of the linear model for viscoelasticity. In this model several Maxwell elements are assembled in parallel. It takes into account that the relaxation does not occur at a single time, but in a set of times. Due to the presence of 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.

Extensional viscosity is a viscosity coefficient when the applied stress is extensional stress. It is often used for characterizing polymer solutions. Extensional viscosity can be measured using rheometers that apply extensional stress. Acoustic rheometer is one example of such devices.

<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.

<span class="mw-page-title-main">Fiber-reinforced composite</span>

A fiber-reinforced composite (FRC) is a composite building material that consists of three components:

  1. the fibers as the discontinuous or dispersed phase,
  2. the matrix as the continuous phase, and
  3. the fine interphase region, also known as the interface.
<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.

<span class="mw-page-title-main">Cinna Lomnitz</span> Mexican researcher

Cinna Lomnitz Aronsfrau was a Chilean-Mexican geophysicist known for his contributions in the fields of rock mechanics and seismology.

Magnetic diffusion refers to the motion of magnetic fields, typically in the presence of a conducting solid or fluid such as a plasma. The motion of magnetic fields is described by the magnetic diffusion equation and is due primarily to induction and diffusion of magnetic fields through the material. The magnetic diffusion equation is a partial differential equation commonly used in physics. Understanding the phenomenon is essential to magnetohydrodynamics and has important consequences in astrophysics, geophysics, and electrical engineering.

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. Holm, Sverre (2024). Acoustic wave equations and four ways media may perturb the speed of sound (PDF). Vol. 1.3. University of Oslo.
  2. 1 2 3 4 5 David Roylance, "Engineering Viscoelasticity" (October 24, 2001) http://ocw.mit.edu/courses/materials-science-and-engineering/3-11-mechanics-of-materials-fall-1999/modules/MIT3_11F99_visco.pdf
  3. Krystyn J. Van Vliet, MIT course 3.032 Lecture, October 23, 2006 http://stellar.mit.edu/S/course/3/fa06/3.032/index.html
  4. Joseph, Daniel D. (2013-11-27). Fluid Dynamics of Viscoelastic Liquids. Springer Science & Business Media. ISBN   9781461244622.