Jump to content

Stokes flow: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
i added it..turns out i am not so sure
No edit summary
Line 1: Line 1:
[[Image:Stokes sphere.svg|thumb|right|200px|An object moving through a gas or liquid experiences a [[force]] in direction opposite to its motion. [[Terminal velocity]] is achieved when the drag force is equal in magnitude but opposite in direction to the force propelling the object. Shown is a [[sphere]] in Stokes flow, at very low [[Reynolds number]].]]
[[Image:Stokes sphere.svg|thumb|right|200px|An object moving through a gas or liquid experiences a [[force]] in direction opposite to its motion. [[Terminal velocity]] is achieved when the drag force is equal in magnitude but opposite in direction to the force propelling the object. Shown is a [[sphere]] in Stokes flow, at very low [[Reynolds number]].]]
'''Stokes flow''' (named after [[George Gabriel Stokes]]), also named '''creeping flow''', is a type of [[fluid flow]] where [[advection|advective]] [[inertia|inertial]] forces are small compared with [[Viscosity|viscous]] forces. The [[Reynolds number]] is low, i.e. <math>\textit{Re} \ll 1</math>. 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 [[microorganism]]s and [[sperm]]<ref>Dusenbery, David B. (2009). ''Living at Micro Scale''. Harvard University Press, Cambridge, Mass. ISBN 978-0-674-03116-6.</ref> and the flow of [[lava]]. In technology, it occurs in [[paint]], [[Microelectromechanical Systems|MEMS]] devices, and in the flow of viscous [[polymer]]s generally.
'''Stokes flow''' (named after [[George Gabriel Stokes]]), also named '''creeping flow''', is a type of [[fluid flow]] where [[advection|advective]] [[inertia|inertial]] forces are small compared with [[Viscosity|viscous]] forces. The [[Reynolds number]] is low, i.e. <math>\textit{Re} \ll 1</math>. 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 [[microorganism]]s and [[sperm]]<ref>Dusenbery, David B. (2009). ''Living at Micro Scale''. Harvard University Press, Cambridge, Mass. ISBN 978-0-674-03116-6.</ref> and the flow of [[lava]]. In technology, it occurs in [[paint]], [[Microelectromechanical Systems|MEMS]] devices, and in the flow of viscous [[polymer]]s generally. The primary singularity of Stokes flow is the Stokeslet, which is associated with a singular point force embedded in a Stokes flow. From its derivatives other fundamental singularities can be obtained <ref name="Chwang">Chwang, A. and Wu, T. (1974). [http://www.nada.kth.se/~annak/Chwang&Wu.pdf "Hydromechanics of low-Reynolds-number flow. Part 2. Singularity method for Stokes flows"]. ''J. Fluid Mech. 62''(6), part 4, 787-815.</ref>.


== Stokes equations ==
== Stokes equations ==

Revision as of 04:14, 7 February 2012

An object moving through a gas or liquid experiences a force in direction opposite to its motion. Terminal velocity is achieved when the drag force is equal in magnitude but opposite in direction to the force propelling the object. Shown is a sphere in Stokes flow, at very low Reynolds number.

Stokes flow (named after George Gabriel Stokes), also named creeping flow, 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[1] and the flow of lava. In technology, it occurs in paint, MEMS devices, and in the flow of viscous polymers generally. The primary singularity of Stokes flow is the Stokeslet, which is associated with a singular point force embedded in a Stokes flow. From its derivatives other fundamental singularities can be obtained [2].

Stokes equations

For this type of flow, the inertial forces are assumed to be negligible and the Navier–Stokes equations simplify to give the Stokes equations:

where is the stress tensor, and an applied body force. There is also an equation for conservation of mass. In the common case of an incompressible Newtonian fluid, the Stokes equations are:

Here is the velocity of the fluid, is the gradient of the pressure, and is the dynamic viscosity.

Properties

The Stokes equations represent a considerable simplification of the full Navier–Stokes equations, especially in the incompressible Newtonian case.[3][4] [5][6]

Instantaneity
A Stokes flow has no dependence on time other than through time-dependent boundary conditions. This means that, given the boundary conditions of a Stokes flow, the flow can be found without knowledge of the flow at any other time.
Time-reversibility
An immediate consequence of instantaneity, time-reversibility means then a time-reversed Stokes flow solves the same equations as the original Stokes flow. This property can sometimes be used (in conjunction with linearity and symmetry in the boundary conditions) to derive results about a flow without solving it fully. Time reversibility means that it is difficult to mix two fluids using creeping flow; a dramatic demonstration is possible of apparently mixing two fluids and then unmixing them by reversing the direction of the mixer.[7] [8]

While these properties are true for incompressible Newtonian Stokes flows, the non-linear and sometimes time-dependent nature of non-Newtonian fluids means that they do not hold in the more general case.

Methods of solution

1. By stream function

The equation for an incompressible Newtonian Stokes flow can be solved by the stream function method in planar or in 3-D axisymmetric cases

Type of function Geometry Equation Comments
Stream function 2-D planar or (biharmonic equation) is the Laplacian operator in two dimensions
Stokes stream function 3-D spherical where For derivation of the operator see Stokes_stream_function#Vorticity
Stokes stream function 3-D cylindrical where For see [9]

.

2. By Green's function

The linearity of the Stokes equations in the case of an incompressible Newtonian fluid means that a Green's function, , for the equations can be found, where is the position vector. The solution for the pressure and velocity due to a point force acting at the origin with and vanishing at infinity is given by[10]

where

  is a second-rank tensor (or more accurately tensor field) known as the Oseen tensor (after Carl Wilhelm Oseen).

For a continuous-force distribution (density) the solution (again vanishing at infinity) can then be constructed by superposition:

3. By Papkovich–Neuber solution

The Papkovich–Neuber solution represents the velocity and pressure fields of an incompressible Newtonian Stokes flow in terms of two harmonic potentials.

4. By boundary element method

Certain problems, such as the evolution of the shape of a bubble in a Stokes flow, are conducive to numerical solution by the boundary element method. This technique can be applied to both 2- and 3-dimensional flows.

See also

References

  1. ^ Dusenbery, David B. (2009). Living at Micro Scale. Harvard University Press, Cambridge, Mass. ISBN 978-0-674-03116-6.
  2. ^ Chwang, A. and Wu, T. (1974). "Hydromechanics of low-Reynolds-number flow. Part 2. Singularity method for Stokes flows". J. Fluid Mech. 62(6), part 4, 787-815.
  3. ^ Leal, L.G. (2007). Advanced Transport Phenomena:Fluid Mechanics and Convective Transport Processes.
  4. ^ Kirby, B.J. (2010). Micro- and Nanoscale Fluid Mechanics: Transport in Microfluidic Devices. Cambridge University Press. ISBN 978-0521119030.
  5. ^ Happel, J. & Brenner, H. (1981) Low Reynolds Number Hydrodynamics, Springer. ISBN 9001371159.
  6. ^ Batchelor, G. (2000). Introduction to Fluid Mechanics.
  7. ^ Dusenbery, David B. (2009). Living at Micro Scale, pp.46. Harvard University Press, Cambridge, Mass. ISBN 978-0-674-03116-6.
  8. ^ http://www.youtube.com/watch?v=p08_KlTKP50&feature=related
  9. ^ Payne, LE (1960). "The Stokes flow problem for a class of axially symmetric bodies". Journal of Fluid Mechanics. 7 (04): 529–549. Bibcode:1960JFM.....7..529P. doi:10.1017/S002211206000027X. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  10. ^ Kim, S. & Karrila, S. J. (2005) Microhydrodynamics: Principles and Selected Applications, Dover. ISBN 0486442195.
  • Ockendon, H. & Ockendon J. R. (1995) Viscous Flow, Cambridge University Press. ISBN 0521458811.