Hele-Shaw flow

Hele-Shaw flow is defined as flow taking place between two parallel flat plates separated by a narrow gap satisfying certain conditions, named after Henry Selby Hele-Shaw, who studied the problem in 1898.[1][2] Various problems in fluid mechanics can be approximated to Hele-Shaw flows and thus the research of these flows is of importance. Approximation to Hele-Shaw flow is specifically important to micro-flows. This is due to manufacturing techniques, which creates shallow planar configurations, and the typically low Reynolds numbers of micro-flows.

The conditions that needs to be satisfied are

h l 1 , U h ν h l 1 {\displaystyle {\frac {h}{l}}\ll 1,\qquad {\frac {Uh}{\nu }}{\frac {h}{l}}\ll 1}

where h {\displaystyle h} is the gap width between the plates, U {\displaystyle U} is the characteristic velocity scale, l {\displaystyle l} is the characteristic length scale in directions parallel to the plate and ν {\displaystyle \nu } is the kinematic viscosity. Specifically, the Reynolds number R e = U h / ν {\displaystyle Re=Uh/\nu } need not always be small, but can be order unity or greater as long as it satisfies the condition R e ( h / l ) 1. {\displaystyle Re(h/l)\ll 1.} In terms of the Reynolds number R e l = U l / ν {\displaystyle Re_{l}=Ul/\nu } based on l {\displaystyle l} , the condition becomes R e l ( h / l ) 2 1. {\displaystyle Re_{l}(h/l)^{2}\ll 1.}

The governing equation of Hele-Shaw flows is identical to that of the inviscid potential flow and to the flow of fluid through a porous medium (Darcy's law). It thus permits visualization of this kind of flow in two dimensions.[3][4][5]

Mathematical formulation of Hele-Shaw flows

A schematic description of a Hele-Shaw configuration.

Let x {\displaystyle x} , y {\displaystyle y} be the directions parallel to the flat plates, and z {\displaystyle z} the perpendicular direction, with h {\displaystyle h} being the gap between the plates (at z = 0 , h {\displaystyle z=0,h} ) and l {\displaystyle l} be the relevant characteristic length scale in the x y {\displaystyle xy} -directions. Under the limits mentioned above, the incompressible Navier–Stokes equations, in the first approximation becomes[6]

p x = μ 2 v x z 2 , p y = μ 2 v y z 2 , p z = 0 , v x x + v y y + v z z = 0 , {\displaystyle {\begin{aligned}{\frac {\partial p}{\partial x}}=\mu {\frac {\partial ^{2}v_{x}}{\partial z^{2}}},\quad {\frac {\partial p}{\partial y}}&=\mu {\frac {\partial ^{2}v_{y}}{\partial z^{2}}},\quad {\frac {\partial p}{\partial z}}=0,\\{\frac {\partial v_{x}}{\partial x}}+{\frac {\partial v_{y}}{\partial y}}+{\frac {\partial v_{z}}{\partial z}}&=0,\\\end{aligned}}}

where μ {\displaystyle \mu } is the viscosity. These equations are similar to boundary layer equations, except that there are no non-linear terms. In the first approximation, we then have, after imposing the non-slip boundary conditions at z = 0 , h {\displaystyle z=0,h} ,

p = p ( x , y ) , v x = 1 2 μ p x z ( h z ) , v y = 1 2 μ p y z ( h z ) {\displaystyle {\begin{aligned}p&=p(x,y),\\v_{x}&=-{\frac {1}{2\mu }}{\frac {\partial p}{\partial x}}z(h-z),\\v_{y}&=-{\frac {1}{2\mu }}{\frac {\partial p}{\partial y}}z(h-z)\end{aligned}}}

The equation for p {\displaystyle p} is obtained from the continuity equation. Integrating the continuity equation from across the channel and imposing no-penetration boundary conditions at the walls, we have

0 h ( v x x + v y y ) d z = 0 , {\displaystyle \int _{0}^{h}\left({\frac {\partial v_{x}}{\partial x}}+{\frac {\partial v_{y}}{\partial y}}\right)dz=0,}

which leads to the Laplace Equation:

2 p x 2 + 2 p y 2 = 0. {\displaystyle {\frac {\partial ^{2}p}{\partial x^{2}}}+{\frac {\partial ^{2}p}{\partial y^{2}}}=0.}

This equation is supplemented by appropriate boundary conditions. For example, no-penetration boundary conditions on the side walls become: p n = 0 {\displaystyle {\mathbf {\nabla } }p\cdot \mathbf {n} =0} , where n {\displaystyle \mathbf {n} } is a unit vector perpendicular to the side wall (note that on the side walls, non-slip boundary conditions cannot be imposed). The boundaries may also be regions exposed to constant pressure in which case a Dirichlet boundary condition for p {\displaystyle p} is appropriate. Similarly, periodic boundary conditions can also be used. It can also be noted that the vertical velocity component in the first approximation is

v z = 0 {\displaystyle v_{z}=0}

that follows from the continuity equation. While the velocity magnitude v x 2 + v y 2 {\displaystyle {\sqrt {v_{x}^{2}+v_{y}^{2}}}} varies in the z {\displaystyle z} direction, the velocity-vector direction tan 1 ( v y / v x ) {\displaystyle \tan ^{-1}(v_{y}/v_{x})} is independent of z {\displaystyle z} direction, that is to say, streamline patterns at each level are similar. The vorticity vector ω {\displaystyle {\boldsymbol {\omega }}} has the components[6]

ω x = 1 2 μ p y ( h 2 z ) , ω y = 1 2 μ p x ( h 2 z ) , ω z = 0. {\displaystyle \omega _{x}={\frac {1}{2\mu }}{\frac {\partial p}{\partial y}}(h-2z),\quad \omega _{y}=-{\frac {1}{2\mu }}{\frac {\partial p}{\partial x}}(h-2z),\quad \omega _{z}=0.}

Since ω z = 0 {\displaystyle \omega _{z}=0} , the streamline patterns in the x y {\displaystyle xy} -plane thus correspond to potential flow (irrotational flow). Unlike potential flow, here the circulation Γ {\displaystyle \Gamma } around any closed contour C {\displaystyle C} (parallel to the x y {\displaystyle xy} -plane), whether it encloses a solid object or not, is zero,

Γ = C v x d x + v y d y = 1 2 μ z ( h z ) C ( p x d x + p y d y ) = 0 {\displaystyle \Gamma =\oint _{C}v_{x}dx+v_{y}dy=-{\frac {1}{2\mu }}z(h-z)\oint _{C}\left({\frac {\partial p}{\partial x}}dx+{\frac {\partial p}{\partial y}}dy\right)=0}

where the last integral is set to zero because p {\displaystyle p} is a single-valued function and the integration is done over a closed contour.

Depth-averaged form

In a Hele-Shaw channel, one can define the depth-averaged version of any physical quantity, say φ {\displaystyle \varphi } by

φ 1 h 0 h φ d z . {\displaystyle \langle \varphi \rangle \equiv {\frac {1}{h}}\int _{0}^{h}\varphi dz.}

Then the two-dimensional depth-averaged velocity vector u v x y {\displaystyle \mathbf {u} \equiv \langle \mathbf {v} _{xy}\rangle } , where v x y = ( v x , v y ) {\displaystyle \mathbf {v} _{xy}=(v_{x},v_{y})} , satisfies the Darcy's law,

12 μ h 2 u = p with u = 0. {\displaystyle -{\frac {12\mu }{h^{2}}}\mathbf {u} =\nabla p\quad {\text{with}}\quad \nabla \cdot \mathbf {u} =0.}

Further, ω = 0. {\displaystyle \langle {\boldsymbol {\omega }}\rangle =0.}

Hele-Shaw cell

The term Hele-Shaw cell is commonly used for cases in which a fluid is injected into the shallow geometry from above or below the geometry, and when the fluid is bounded by another liquid or gas.[7] For such flows the boundary conditions are defined by pressures and surface tensions.

See also

  • Diffusion-limited aggregation
  • Lubrication theory
  • Thin-film equation
  • Hele-Shaw clutch

References

  1. ^ Shaw, Henry S. H. (1898). Investigation of the nature of surface resistance of water and of stream-line motion under certain experimental conditions. Inst. N.A. OCLC 17929897.[page needed]
  2. ^ Hele-Shaw, H. S. (1 May 1898). "The Flow of Water". Nature. 58 (1489): 34–36. Bibcode:1898Natur..58...34H. doi:10.1038/058034a0.
  3. ^ Hermann Schlichting,Boundary Layer Theory, 7th ed. New York: McGraw-Hill, 1979.[page needed]
  4. ^ L. M. Milne-Thomson (1996). Theoretical Hydrodynamics. Dover Publications, Inc.
  5. ^ Horace Lamb, Hydrodynamics (1934).[page needed]
  6. ^ a b Acheson, D. J. (1991). Elementary fluid dynamics.
  7. ^ Saffman, P. G. (21 April 2006). "Viscous fingering in Hele-Shaw cells" (PDF). Journal of Fluid Mechanics. 173: 73–94. doi:10.1017/s0022112086001088. S2CID 17003612.