Strain-rate tensor

Last updated
A two-dimensional flow that, at the highlighted point, has only a strain rate component, with no mean velocity or rotational component. Vel decomp field01 term6 s.png
A two-dimensional flow that, at the highlighted point, has only a strain rate component, with no mean velocity or rotational component.

In continuum mechanics, the strain-rate tensor or rate-of-strain tensor is a physical quantity that describes the rate of change of the strain (i.e., the relative deformation) of a material in the neighborhood of a certain point, at a certain moment of time. It can be defined as the derivative of the strain tensor with respect to time, or as the symmetric component of the Jacobian matrix (derivative with respect to position) of the flow velocity. In fluid mechanics it also can be described as the velocity gradient, a measure of how the velocity of a fluid changes between different points within the fluid. [1] Though the term can refer to a velocity profile (variation in velocity across layers of flow in a pipe), [2] it is often used to mean the gradient of a flow's velocity with respect to its coordinates. [3] The concept has implications in a variety of areas of physics and engineering, including magnetohydrodynamics, mining and water treatment. [4] [5] [6]

Contents

The strain rate tensor is a purely kinematic concept that describes the macroscopic motion of the material. Therefore, it does not depend on the nature of the material, or on the forces and stresses that may be acting on it; and it applies to any continuous medium, whether solid, liquid or gas.

On the other hand, for any fluid except superfluids, any gradual change in its deformation (i.e. a non-zero strain rate tensor) gives rise to viscous forces in its interior, due to friction between adjacent fluid elements, that tend to oppose that change. At any point in the fluid, these stresses can be described by a viscous stress tensor that is, almost always, completely determined by the strain rate tensor and by certain intrinsic properties of the fluid at that point. Viscous stress also occur in solids, in addition to the elastic stress observed in static deformation; when it is too large to be ignored, the material is said to be viscoelastic.

Dimensional analysis

By performing dimensional analysis, the dimensions of velocity gradient can be determined. The dimensions of velocity are , and the dimensions of distance are . Since the velocity gradient can be expressed as . Therefore, the velocity gradient has the same dimensions as this ratio, i.e., .

In continuum mechanics

In 3 dimensions, the gradient of the velocity is a second-order tensor which can be expressed as the matrix :

can be decomposed into the sum of a symmetric matrix and a skew-symmetric matrix as follows

is called the strain rate tensor and describes the rate of stretching and shearing. is called the spin tensor and describes the rate of rotation. [7]

Relationship between shear stress and the velocity field

Sir Isaac Newton proposed that shear stress is directly proportional to the velocity gradient: [8]

The constant of proportionality, , is called the dynamic viscosity.

Formal definition

Consider a material body, solid or fluid, that is flowing and/or moving in space. Let v be the velocity field within the body; that is, a smooth function from R3 × R such that v(p, t) is the macroscopic velocity of the material that is passing through the point p at time t.

The velocity v(p + r, t) at a point displaced from p by a small vector r can be written as a Taylor series:

where v the gradient of the velocity field, understood as a linear map that takes a displacement vector r to the corresponding change in the velocity.

Vel decomp field01 term0 s.png
Total field v(p + r).
Vel decomp field01 term3 s.png
Constant part v(p).
Vel decomp field01 term4 s.png
Linear part (∇v)(p, t)(r).
Vel decomp field01 term2 s.png
Non-linear residual.
The velocity field v(p + r, t) of an arbitrary flow around a point p (red dot), at some instant t, and the terms of its first-order Taylor approximation about p. The third component of the velocity (out of the screen) is assumed to be zero everywhere.

In an arbitrary reference frame, v is related to the Jacobian matrix of the field, namely in 3 dimensions it is the 3 × 3 matrix

where vi is the component of v parallel to axis i and jf denotes the partial derivative of a function f with respect to the space coordinate xj. Note that J is a function of p and t.

In this coordinate system, the Taylor approximation for the velocity near p is

or simply

if v and r are viewed as 3 × 1 matrices.

Symmetric and antisymmetric parts

Vel decomp field01 term6 s.png
The symmetric part E(p, t)(r) (strain rate) of the linear term of the example flow.
Vel decomp field01 term5 s.png
The antisymmetric part R(p, t)(r) (rotation) of the linear term.

Any matrix can be decomposed into the sum of a symmetric matrix and an antisymmetric matrix. Applying this to the Jacobian matrix with symmetric and antisymmetric components E and R respectively:

This decomposition is independent of coordinate system, and so has physical significance. Then the velocity field may be approximated as

that is,

The antisymmetric term R represents a rigid-like rotation of the fluid about the point p. Its angular velocity is

The product ∇ × v is called the vorticity of the vector field. A rigid rotation does not change the relative positions of the fluid elements, so the antisymmetric term R of the velocity gradient does not contribute to the rate of change of the deformation. The actual strain rate is therefore described by the symmetric E term, which is the strain rate tensor.

Shear rate and compression rate

Vel decomp field01 term7 s.png
The spherical part S(p, t)(r) (uniform expansion, or compression, rate) of the strain rate tensor E(p, t)(r).
Vel decomp field01 term8 s.png
The deviatoric part D(p, t)(r) (shear rate) of the strain rate tensor E(p, t)(r).

The symmetric term E (the rate-of-strain tensor) can be broken down further as the sum of a scalar times the unit tensor, that represents a gradual isotropic expansion or contraction; and a traceless symmetric tensor which represents a gradual shearing deformation, with no change in volume: [9]

That is,

Here δ is the unit tensor, such that δij is 1 if i = j and 0 if ij. This decomposition is independent of the choice of coordinate system, and is therefore physically significant.

The trace of the expansion rate tensor is the divergence of the velocity field:

which is the rate at which the volume of a fixed amount of fluid increases at that point.

The shear rate tensor is represented by a symmetric 3 × 3 matrix, and describes a flow that combines compression and expansion flows along three orthogonal axes, such that there is no change in volume. This type of flow occurs, for example, when a rubber strip is stretched by pulling at the ends, or when honey falls from a spoon as a smooth unbroken stream.

For a two-dimensional flow, the divergence of v has only two terms and quantifies the change in area rather than volume. The factor 1/3 in the expansion rate term should be replaced by 1/2 in that case.

Examples

The study of velocity gradients is useful in analysing path dependent materials and in the subsequent study of stresses and strains; e.g., Plastic deformation of metals. [3] The near-wall velocity gradient of the unburned reactants flowing from a tube is a key parameter for characterising flame stability. [5] :1–3 The velocity gradient of a plasma can define conditions for the solutions to fundamental equations in magnetohydrodynamics. [4]

Fluid in a pipe

Consider the velocity field of a fluid flowing through a pipe. The layer of fluid in contact with the pipe tends to be at rest with respect to the pipe. This is called the no slip condition. [10] If the velocity difference between fluid layers at the centre of the pipe and at the sides of the pipe is sufficiently small, then the fluid flow is observed in the form of continuous layers. This type of flow is called laminar flow.

The flow velocity difference between adjacent layers can be measured in terms of a velocity gradient, given by . Where is the difference in flow velocity between the two layers and is the distance between the layers.

See also

Related Research Articles

Continuum mechanics is a branch of mechanics that deals with the deformation of and transmission of forces through materials modeled as a continuous medium rather than as discrete particles. The French mathematician Augustin-Louis Cauchy was the first to formulate such models in the 19th century.

<span class="mw-page-title-main">Divergence</span> Vector operator in vector calculus

In vector calculus, divergence is a vector operator that operates on a vector field, producing a scalar field giving the quantity of the vector field's source at each point. More technically, the divergence represents the volume density of the outward flux of a vector field from an infinitesimal volume around a given point.

<span class="mw-page-title-main">Gradient</span> Multivariate derivative (mathematics)

In vector calculus, the gradient of a scalar-valued differentiable function of several variables is the vector field whose value at a point gives the direction and the rate of fastest increase. The gradient transforms like a vector under change of basis of the space of variables of . If the gradient of a function is non-zero at a point , the direction of the gradient is the direction in which the function increases most quickly from , and the magnitude of the gradient is the rate of increase in that direction, the greatest absolute directional derivative. Further, a point where the gradient is the zero vector is known as a stationary point. The gradient thus plays a fundamental role in optimization theory, where it is used to minimize a function by gradient descent. In coordinate-free terms, the gradient of a function may be defined by:

<span class="mw-page-title-main">Navier–Stokes equations</span> Equations describing the motion of viscous fluid substances

The Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances. They were named after French engineer and physicist Claude-Louis Navier and the Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842–1850 (Stokes).

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">Stream function</span> Function for incompressible divergence-free flows in two dimensions

In fluid dynamics, two types of stream function are defined:

<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.

A Newtonian fluid is a fluid in which the viscous stresses arising from its flow are at every point linearly correlated to the local strain rate — the rate of change of its deformation over time. Stresses are proportional to the rate of change of the fluid's velocity vector.

<span class="mw-page-title-main">Simple shear</span> Translation which preserves parallelism

Simple shear is a deformation in which parallel planes in a material remain parallel and maintain a constant distance, while translating relative to each other.

In continuum mechanics, the finite strain theory—also called large strain theory, or large deformation theory—deals with deformations in which strains and/or rotations are large enough to invalidate assumptions inherent in infinitesimal strain theory. In this case, the undeformed and deformed configurations of the continuum are significantly different, requiring a clear distinction between them. This is commonly the case with elastomers, plastically deforming materials and other fluids and biological soft tissue.

In continuum mechanics, including fluid dynamics, an upper-convected time derivative or Oldroyd derivative, named after James G. Oldroyd, is the rate of change of some tensor property of a small parcel of fluid that is written in the coordinate system rotating and stretching with the fluid.

The derivation of the Navier–Stokes equations as well as its application and formulation for different families of fluids, is an important exercise in fluid dynamics with applications in mechanical engineering, physics, chemistry, heat transfer, and electrical engineering. A proof explaining the properties and bounds of the equations, such as Navier–Stokes existence and smoothness, is one of the important unsolved problems in mathematics.

In continuum mechanics the flow velocity in fluid dynamics, also macroscopic velocity in statistical mechanics, or drift velocity in electromagnetism, is a vector field used to mathematically describe the motion of a continuum. The length of the flow velocity vector is scalar, the flow speed. It is also called velocity field; when evaluated along a line, it is called a velocity profile.

<span class="mw-page-title-main">Diffusion</span> Transport of dissolved species from the highest to the lowest concentration region

Diffusion is the net movement of anything generally from a region of higher concentration to a region of lower concentration. Diffusion is driven by a gradient in Gibbs free energy or chemical potential. It is possible to diffuse "uphill" from a region of lower concentration to a region of higher concentration, as in spinodal decomposition. Diffusion is a stochastic process due to the inherent randomness of the diffusing entity and can be used to model many real-life stochastic scenarios. Therefore, diffusion and the corresponding mathematical models are used in several fields beyond physics, such as statistics, probability theory, information theory, neural networks, finance, and marketing.

In continuum mechanics, a compatible deformation tensor field in a body is that unique tensor field that is obtained when the body is subjected to a continuous, single-valued, displacement field. Compatibility is the study of the conditions under which such a displacement field can be guaranteed. Compatibility conditions are particular cases of integrability conditions and were first derived for linear elasticity by Barré de Saint-Venant in 1864 and proved rigorously by Beltrami in 1886.

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 word.

<span class="mw-page-title-main">Objective stress rate</span>

In continuum mechanics, objective stress rates are time derivatives of stress that do not depend on the frame of reference. Many constitutive equations are designed in the form of a relation between a stress-rate and a strain-rate. The mechanical response of a material should not depend on the frame of reference. In other words, material constitutive equations should be frame-indifferent (objective). If the stress and strain measures are material quantities then objectivity is automatically satisfied. However, if the quantities are spatial, then the objectivity of the stress-rate is not guaranteed even if the strain-rate is objective.

In fluid mechanics, Helmholtz minimum dissipation theorem states that the steady Stokes flow motion of an incompressible fluid has the smallest rate of dissipation than any other incompressible motion with the same velocity on the boundary. The theorem also has been studied by Diederik Korteweg in 1883 and by Lord Rayleigh in 1913.

The streamline upwind Petrov–Galerkin pressure-stabilizing Petrov–Galerkin formulation for incompressible Navier–Stokes equations can be used for finite element computations of high Reynolds number incompressible flow using equal order of finite element space by introducing additional stabilization terms in the Navier–Stokes Galerkin formulation.

References

  1. Carl Schaschke (2014). A Dictionary of Chemical Engineering. Oxford University Press. ISBN   9780199651450.
  2. "Infoplease: Viscosity: The Velocity Gradient".
  3. 1 2 "Velocity gradient at continuummechanics.org".
  4. 1 2 Zhang, Zujin (June 2017), "Generalized MHD System with Velocity Gradient in Besov Spaces of Negative Order", Acta Applicandae Mathematicae, 149 (1): 139–144, doi:10.1007/s10440-016-0091-0, ISSN   1572-9036, S2CID   207075598
  5. 1 2 Grumer, J.; Harris, M. E.; Rowe, V. R. (Jul 1956), Fundamental Flashback, Blowoff, and Yellow-Tip Limits of Fuel Gas-Air Mixtures (PDF), Bureau of Mines
  6. Rojas, J.C.; Moreno, B.; Garralón, G.; Plaza, F.; Pérez, J.; Gómez, M.A. (2010), "Influence of velocity gradient in a hydraulic flocculator on NOM removal by aerated spiral-wound ultrafiltration membranes (ASWUF)", Journal of Hazardous Materials, 178 (1): 535–540, doi:10.1016/j.jhazmat.2010.01.116, ISSN   0304-3894, PMID   20153578
  7. Gonzalez, O.; Stuart, A. M. (2008). A First Course in Continuum Mechanics. Cambridge Texts in Applied Mathematics. Cambridge University Press. pp. 134–135.
  8. Batchelor, G.K. (2000). An Introduction to Fluid Dynamics. Cambridge Mathematical Library. Cambridge University Press. p. 145. ISBN   9780521663960.
  9. Landau, L. D.; Lifshitz, E. M. (1997). Fluid Mechanics. Translated by Sykes, J. B.; Reid, W. H. (2nd ed.). Butterworth Heinemann. ISBN   0-7506-2767-0.
  10. Levicky, R. "Review of fluid mechanics terminology" (PDF).