Couette flow

Last updated

In fluid dynamics, Couette flow is the flow of a viscous fluid in the space between two surfaces, one of which is moving tangentially relative to the other. The relative motion of the surfaces imposes a shear stress on the fluid and induces flow. Depending on the definition of the term, there may also be an applied pressure gradient in the flow direction.

Contents

The Couette configuration models certain practical problems, like the Earth's mantle and atmosphere, [1] and flow in lightly loaded journal bearings. It is also employed in viscometry and to demonstrate approximations of reversibility. [2] [3]

It is named after Maurice Couette, a Professor of Physics at the French University of Angers in the late 19th century.

Planar Couette flow

Simple Couette configuration using two infinite flat plates. Laminar shear.svg
Simple Couette configuration using two infinite flat plates.

Couette flow is frequently used in undergraduate physics and engineering courses to illustrate shear-driven fluid motion. A simple configuration corresponds to two infinite, parallel plates separated by a distance ; one plate translates with a constant relative velocity in its own plane. Neglecting pressure gradients, the Navier–Stokes equations simplify to

where is the spatial coordinate normal to the plates and is the velocity field. This equation reflects the assumption that the flow is unidirectional — that is, only one of the three velocity components is non-trivial. If the lower plate corresponds to , the boundary conditions are and . The exact solution

can be found by integrating twice and solving for the constants using the boundary conditions. A notable aspect of the flow is that shear stress is constant throughout the domain. In particular, the first derivative of the velocity, , is constant. According to Newton's Law of Viscosity (Newtonian fluid), the shear stress is the product of this expression and the (constant) fluid viscosity.

Startup

Startup Couette flow StartupCouette.pdf
Startup Couette flow

In reality, the Couette solution is not reached instantaneously. The "startup problem" describing the approach to steady state is given by

subject to the initial condition

and with the same boundary conditions as the steady flow:

The problem can be made homogeneous by subtracting the steady solution. Then, applying separation of variables leads to the solution: [4]

.

The timescale describing relaxation to steady state is , as illustrated in the figure. The time required to reach the steady state depends only on the spacing between the plates and the kinematic viscosity of the fluid, but not on .

Planar flow with pressure gradient

A more general Couette flow includes a constant pressure gradient in a direction parallel to the plates. The Navier–Stokes equations are

where is the dynamic viscosity. Integrating the above equation twice and applying the boundary conditions (same as in the case of Couette flow without pressure gradient) gives

The pressure gradient can be positive (adverse pressure gradient) or negative (favorable pressure gradient). In the limiting case of stationary plates (), the flow is referred to as Plane Poiseuille flow, and has a symmetric (with reference to the horizontal mid-plane) parabolic velocity profile. [5]

Compressible flow

Compressible Couette flow for
M
=
0
{\displaystyle \mathrm {M} =0} CompCouette.pdf
Compressible Couette flow for
Compressible Couette flow for
M
2
P
r
=
7.5
{\displaystyle \mathrm {M} ^{2}\mathrm {Pr} =7.5} CompCouette2.pdf
Compressible Couette flow for

In incompressible flow, the velocity profile is linear because the fluid temperature is constant. When the upper and lower walls are maintained at different temperatures, the velocity profile is more complicated. However, it has an exact implicit solution as shown by C. R. Illingworth in 1950. [6]

Consider the plane Couette flow with lower wall at rest and the upper wall in motion with constant velocity . Denote fluid properties at the lower wall with subscript and properties at the upper wall with subscript . The properties and the pressure at the upper wall are prescribed and taken as reference quantities. Let be the distance between the two walls. The boundary conditions are

where is the specific enthalpy and is the specific heat. Conservation of mass and -momentum requires everywhere in the flow domain. Conservation of energy and -momentum reduce to

where is the wall shear stress. The flow does not depend on the Reynolds number , but rather on the Prandtl number and the Mach number , where is the thermal conductivity, is the speed of sound and is the specific heat ratio. Introduce the non-dimensional variables

In terms of these quantities, the solutions are

where is the heat transferred per unit time per unit area from the lower wall. Thus are implicit functions of . One can also write the solution in terms of the recovery temperature and recovery enthalpy evaluated at the temperature of an insulated wall i.e., the values of and for which .[ clarification needed ] Then the solution is

If the specific heat is constant, then . When and , then and are constant everywhere, thus recovering the incompressible Couette flow solution. Otherwise, one must know the full temperature dependence of . While there is no simple expression for that is both accurate and general, there are several approximations for certain materials — see, e.g., temperature dependence of viscosity. When and , the recovery quantities become unity . For air, the values are commonly used, and the results for this case are shown in the figure.

The effects of dissociation and ionization (i.e., is not constant) have also been studied; in that case the recovery temperature is reduced by the dissociation of molecules. [7]

Rectangular channel

Couette flow for square channel Couetter.pdf
Couette flow for square channel
Couette flow with h/l=0.1 Couetter1.pdf
Couette flow with h/l=0.1

One-dimensional flow is valid when both plates are infinitely long in the streamwise () and spanwise () directions. When the spanwise length is finite, the flow becomes two-dimensional and is a function of both and . However, the infinite length in the streamwise direction must be retained in order to ensure the unidirectional nature of the flow.

As an example, consider an infinitely long rectangular channel with transverse height and spanwise width , subject to the condition that the top wall moves with a constant velocity . Without an imposed pressure gradient, the Navier–Stokes equations reduce to

with boundary conditions

Using separation of variables, the solution is given by

When , the planar Couette flow is recovered, as shown in the figure.

Coaxial cylinders

Taylor–Couette flow is a flow between two rotating, infinitely long, coaxial cylinders. [8] The original problem was solved by Stokes in 1845, [9] but Geoffrey Ingram Taylor's name was attached to the flow because he studied its stability in a famous 1923 paper. [10]

The problem can be solved in cylindrical coordinates . Denote the radii of the inner and outer cylinders as and . Assuming the cylinders rotate at constant angular velocities and , then the velocity in the -direction is [11]

This equation shows that the effects of curvature no longer allow for constant shear in the flow domain.

Coaxial cylinders of finite length

The classical Taylor–Couette flow problem assumes infinitely long cylinders; if the cylinders have non-negligible finite length , then the analysis must be modified (though the flow is still unidirectional). For , the finite-length problem can be solved using separation of variables or integral transforms, giving: [12]

where are the Modified Bessel functions of the first and second kind.

See also

Related Research Articles

<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 thermal fluid dynamics, the Nusselt number is the ratio of convective to conductive heat transfer at a boundary in a fluid. Convection includes both advection and diffusion (conduction). The conductive component is measured under the same conditions as the convective but for a hypothetically motionless fluid. It is a dimensionless number, closely related to the fluid's Rayleigh number.

In fluid mechanics, the Grashof number is a dimensionless number which approximates the ratio of the buoyancy to viscous forces acting on a fluid. It frequently arises in the study of situations involving natural convection and is analogous to the Reynolds number.

<span class="mw-page-title-main">Fokker–Planck equation</span> Partial differential equation

In statistical mechanics and information theory, the Fokker–Planck equation is a partial differential equation that describes the time evolution of the probability density function of the velocity of a particle under the influence of drag forces and random forces, as in Brownian motion. The equation can be generalized to other observables as well. The Fokker-Planck equation has multiple applications in information theory, graph theory, data science, finance, economics etc.

In mathematical analysis, Hölder's inequality, named after Otto Hölder, is a fundamental inequality between integrals and an indispensable tool for the study of Lp spaces.

<span class="mw-page-title-main">Helmholtz free energy</span> Thermodynamic potential

In thermodynamics, the Helmholtz free energy is a thermodynamic potential that measures the useful work obtainable from a closed thermodynamic system at a constant temperature (isothermal). The change in the Helmholtz energy during a process is equal to the maximum amount of work that the system can perform in a thermodynamic process in which temperature is held constant. At constant temperature, the Helmholtz free energy is minimized at equilibrium.

In physics, a wave vector is a vector used in describing a wave, with a typical unit being cycle per metre. It has a magnitude and direction. Its magnitude is the wavenumber of the wave, and its direction is perpendicular to the wavefront. In isotropic media, this is also the direction of wave propagation.

<span class="mw-page-title-main">Stokes flow</span> Type of fluid flow

Stokes flow, also named creeping flow or creeping motion, is a type of fluid flow where advective inertial forces are small compared with viscous forces. The Reynolds number is low, i.e. . This is a typical situation in flows where the fluid velocities are very slow, the viscosities are very large, or the length-scales of the flow are very small. Creeping flow was first studied to understand lubrication. In nature, this type of flow occurs in the swimming of microorganisms and sperm. In technology, it occurs in paint, MEMS devices, and in the flow of viscous polymers generally.

<span class="mw-page-title-main">Euler–Bernoulli beam theory</span> Method for load calculation in construction

Euler–Bernoulli beam theory is a simplification of the linear theory of elasticity which provides a means of calculating the load-carrying and deflection characteristics of beams. It covers the case corresponding to small deflections of a beam that is subjected to lateral loads only. By ignoring the effects of shear deformation and rotatory inertia, it is thus a special case of Timoshenko–Ehrenfest beam theory. It was first enunciated circa 1750, but was not applied on a large scale until the development of the Eiffel Tower and the Ferris wheel in the late 19th century. Following these successful demonstrations, it quickly became a cornerstone of engineering and an enabler of the Second Industrial Revolution.

<span class="mw-page-title-main">Taylor–Couette flow</span>

In fluid dynamics, the Taylor–Couette flow consists of a viscous fluid confined in the gap between two rotating cylinders. For low angular velocities, measured by the Reynolds number Re, the flow is steady and purely azimuthal. This basic state is known as circular Couette flow, after Maurice Marie Alfred Couette, who used this experimental device as a means to measure viscosity. Sir Geoffrey Ingram Taylor investigated the stability of Couette flow in a ground-breaking paper. Taylor's paper became a cornerstone in the development of hydrodynamic stability theory and demonstrated that the no-slip condition, which was in dispute by the scientific community at the time, was the correct boundary condition for viscous flows at a solid boundary.

<span class="mw-page-title-main">Lemniscate elliptic functions</span> Mathematical functions

In mathematics, the lemniscate elliptic functions are elliptic functions related to the arc length of the lemniscate of Bernoulli. They were first studied by Giulio Fagnano in 1718 and later by Leonhard Euler and Carl Friedrich Gauss, among others.

In physics and fluid mechanics, a Blasius boundary layer describes the steady two-dimensional laminar boundary layer that forms on a semi-infinite plate which is held parallel to a constant unidirectional flow. Falkner and Skan later generalized Blasius' solution to wedge flow, i.e. flows in which the plate is not parallel to the flow.

A ratio distribution is a probability distribution constructed as the distribution of the ratio of random variables having two other known distributions. Given two random variables X and Y, the distribution of the random variable Z that is formed as the ratio Z = X/Y is a ratio distribution.

In mathematics, the spectral theory of ordinary differential equations is the part of spectral theory concerned with the determination of the spectrum and eigenfunction expansion associated with a linear ordinary differential equation. In his dissertation, Hermann Weyl generalized the classical Sturm–Liouville theory on a finite closed interval to second order differential operators with singularities at the endpoints of the interval, possibly semi-infinite or infinite. Unlike the classical case, the spectrum may no longer consist of just a countable set of eigenvalues, but may also contain a continuous part. In this case the eigenfunction expansion involves an integral over the continuous part with respect to a spectral measure, given by the Titchmarsh–Kodaira formula. The theory was put in its final simplified form for singular differential equations of even degree by Kodaira and others, using von Neumann's spectral theorem. It has had important applications in quantum mechanics, operator theory and harmonic analysis on semisimple Lie groups.

In nonideal fluid dynamics, the Hagen–Poiseuille equation, also known as the Hagen–Poiseuille law, Poiseuille law or Poiseuille equation, is a physical law that gives the pressure drop in an incompressible and Newtonian fluid in laminar flow flowing through a long cylindrical pipe of constant cross section. It can be successfully applied to air flow in lung alveoli, or the flow through a drinking straw or through a hypodermic needle. It was experimentally derived independently by Jean Léonard Marie Poiseuille in 1838 and Gotthilf Heinrich Ludwig Hagen, and published by Poiseuille in 1840–41 and 1846. The theoretical justification of the Poiseuille law was given by George Stokes in 1845.

A Sommerfeld expansion is an approximation method developed by Arnold Sommerfeld for a certain class of integrals which are common in condensed matter and statistical physics. Physically, the integrals represent statistical averages using the Fermi–Dirac distribution.

The table of chords, created by the Greek astronomer, geometer, and geographer Ptolemy in Egypt during the 2nd century AD, is a trigonometric table in Book I, chapter 11 of Ptolemy's Almagest, a treatise on mathematical astronomy. It is essentially equivalent to a table of values of the sine function. It was the earliest trigonometric table extensive enough for many practical purposes, including those of astronomy. Since the 8th and 9th centuries, the sine and other trigonometric functions have been used in Islamic mathematics and astronomy, reforming the production of sine tables. Khwarizmi and Habash al-Hasib later produced a set of trigonometric tables.

In fluid dynamics, a flow with periodic variations is known as pulsatile flow, or as Womersley flow. The flow profiles was first derived by John R. Womersley (1907–1958) in his work with blood flow in arteries. The cardiovascular system of chordate animals is a very good example where pulsatile flow is found, but pulsatile flow is also observed in engines and hydraulic systems, as a result of rotating mechanisms pumping the fluid.

<span class="mw-page-title-main">Falkner–Skan boundary layer</span> Boundary Layer

In fluid dynamics, the Falkner–Skan boundary layer describes the steady two-dimensional laminar boundary layer that forms on a wedge, i.e. flows in which the plate is not parallel to the flow. It is also representative of flow on a flat plate with an imposed pressure gradient along the plate length, a situation often encountered in wind tunnel flow. It is a generalization of the flat plate Blasius boundary layer in which the pressure gradient along the plate is zero.

<span class="mw-page-title-main">Stokes problem</span>

In fluid dynamics, Stokes problem also known as Stokes second problem or sometimes referred to as Stokes boundary layer or Oscillating boundary layer is a problem of determining the flow created by an oscillating solid surface, named after Sir George Stokes. This is considered one of the simplest unsteady problems that has an exact solution for the Navier-Stokes equations. In turbulent flow, this is still named a Stokes boundary layer, but now one has to rely on experiments, numerical simulations or approximate methods in order to obtain useful information on the flow.

References

  1. Zhilenko et al. (2018)
  2. Guyon et al. (2001), p. 136
  3. Heller (1960)
  4. Pozrikidis (2011), pp. 338–339
  5. Kundu et al. (2016), p. 415
  6. Lagerstrom (1996)
  7. Liepmann et al. (1956, 1957)
  8. Landau and Lifshitz (1987)
  9. Stokes (1845)
  10. Taylor (1923)
  11. Guyon et al. (2001), pp. 163–166
  12. Wendl (1999)

Sources