Thin-film equation

Last updated

In fluid mechanics, the thin-film equation is a partial differential equation that approximately predicts the time evolution of the thickness h of a liquid film that lies on a surface. The equation is derived via lubrication theory which is based on the assumption that the length-scales in the surface directions are significantly larger than in the direction normal to the surface. In the non-dimensional form of the Navier-Stokes equation the requirement is that terms of order ε2 and ε2Re are negligible, where ε ≪ 1 is the aspect ratio and Re is the Reynolds number. This significantly simplifies the governing equations. However, lubrication theory, as the name suggests, is typically derived for flow between two solid surfaces, hence the liquid forms a lubricating layer. The thin-film equation holds when there is a single free surface. With two free surfaces, the flow must be treated as a viscous sheet. [1] [2]

Contents

Definition

The basic form of a 2-dimensional thin film equation is [3] [4] [5]

where the fluid flux is

,

and μ is the viscosity (or dynamic viscosity) of the liquid, h(x,y,t) is film thickness, γ is the interfacial tension between the liquid and the gas phase above it, is the liquid density and the surface shear. The surface shear could be caused by flow of the overlying gas or surface tension gradients. [6] [7] The vectors represent the unit vector in the surface co-ordinate directions, the dot product serving to identify the gravity component in each direction. The vector is the unit vector perpendicular to the surface.

A generalised thin film equation is discussed in [5]

.

When this may represent flow with slip at the solid surface whole describes the thickness of a thin bridge between two masses of fluid in a Hele-Shaw cell. [8] The value represents surface tension driven flow.

A form frequently investigated with regard to the rupture of thin liquid films involves the addition of a disjoining pressure Π(h) in the equation, [9] as in

where the function Π(h) is usually very small in value for moderate-large film thicknesses h and grows very rapidly when h goes very close to zero.

Properties

Physical applications, properties and solution behaviour of the thin-film equation are reviewed in. [3] [5] With the inclusion of phase change at the substrate a form of thin film equation for an arbitrary surface is derived in. [10] A detailed study of the steady-flow of a thin film near a moving contact line is given in. [11] For a yield-stress fluid flow driven by gravity and surface tension is investigated in. [12]

For purely surface tension driven flow it is easy to see that one static (time-independent) solution is a paraboloid of revolution

and this is consistent with the experimentally observed spherical cap shape of a static sessile drop, as a "flat" spherical cap that has small height can be accurately approximated in second order with a paraboloid. This, however, does not handle correctly the circumference of the droplet where the value of the function h(x,y) drops to zero and below, as a real physical liquid film can't have a negative thickness. This is one reason why the disjoining pressure term Π(h) is important in the theory.

One possible realistic form of the disjoining pressure term is [9]

where B, h*, m and n are some parameters. These constants and the surface tension can be approximately related to the equilibrium liquid-solid contact angle through the equation [9] [13]

.

The thin film equation can be used to simulate several behaviors of liquids, such as the fingering instability in gravity driven flow. [14]

The lack of a second-order time derivative in the thin-film equation is a result of the assumption of small Reynold's number in its derivation, which allows the ignoring of inertial terms dependent on fluid density . [14] This is somewhat similar to the situation with Washburn's equation, which describes the capillarity-driven flow of a liquid in a thin tube.

See also

Related Research Articles

<span class="mw-page-title-main">Fick's laws of diffusion</span> Mathematical descriptions of molecular diffusion

Fick's laws of diffusion describe diffusion and were derived by Adolf Fick in 1855. They can be used to solve for the diffusion coefficient, D. Fick's first law can be used to derive his second law which in turn is identical to the diffusion equation.

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

In physics, the Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances, named after French engineer and physicist Claude-Louis Navier and Anglo-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 1851, George Gabriel Stokes derived an expression, now known as Stokes' law, for the frictional force – also called drag force – exerted on spherical objects with very small Reynolds numbers in a viscous fluid. Stokes' law is derived by solving the Stokes flow limit for small Reynolds numbers of the Navier–Stokes equations.

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.

A continuity equation or transport equation is an equation that describes the transport of some quantity. It is particularly simple and powerful when applied to a conserved quantity, but it can be generalized to apply to any extensive quantity. Since mass, energy, momentum, electric charge and other natural quantities are conserved under their respective appropriate conditions, a variety of physical phenomena may be described using continuity equations.

<span class="mw-page-title-main">Scalar potential</span> When potential energy difference depends only on displacement

In mathematical physics, scalar potential, simply stated, describes the situation where the difference in the potential energies of an object in two different positions depends only on the positions, not upon the path taken by the object in traveling from one position to the other. It is a scalar field in three-space: a directionless value (scalar) that depends only on its location. A familiar example is potential energy due to gravity.

<span class="mw-page-title-main">Onsager reciprocal relations</span> Relations between flows and forces, or gradients, in thermodynamic systems

In thermodynamics, the Onsager reciprocal relations express the equality of certain ratios between flows and forces in thermodynamic systems out of equilibrium, but where a notion of local equilibrium exists.

Geometrical optics, or ray optics, is a model of optics that describes light propagation in terms of rays. The ray in geometrical optics is an abstraction useful for approximating the paths along which light propagates under certain circumstances.

In differential geometry, the four-gradient is the four-vector analogue of the gradient from vector calculus.

<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.

In quantum mechanics, the Pauli equation or Schrödinger–Pauli equation is the formulation of the Schrödinger equation for spin-½ particles, which takes into account the interaction of the particle's spin with an external electromagnetic field. It is the non-relativistic limit of the Dirac equation and can be used where particles are moving at speeds much less than the speed of light, so that relativistic effects can be neglected. It was formulated by Wolfgang Pauli in 1927.

<span class="mw-page-title-main">Lubrication theory</span> Flow of fluids within extremely thin regions

In fluid dynamics, lubrication theory describes the flow of fluids in a geometry in which one dimension is significantly smaller than the others. An example is the flow above air hockey tables, where the thickness of the air layer beneath the puck is much smaller than the dimensions of the puck itself.

The intent of this article is to highlight the important points of the derivation of the Navier–Stokes equations as well as its application and formulation for different families of fluids.

In fluid mechanics and mathematics, a capillary surface is a surface that represents the interface between two different fluids. As a consequence of being a surface, a capillary surface has no thickness in slight contrast with most real fluid interfaces.

The Madelung equations, or the equations of quantum hydrodynamics, are Erwin Madelung's equivalent alternative formulation of the Schrödinger equation, written in terms of hydrodynamical variables, similar to the Navier–Stokes equations of fluid dynamics. The derivation of the Madelung equations is similar to the de Broglie–Bohm formulation, which represents the Schrödinger equation as a quantum Hamilton–Jacobi equation.

Hele-Shaw flow is defined as Stokes flow between two parallel flat plates separated by an infinitesimally small gap, named after Henry Selby Hele-Shaw, who studied the problem in 1898. 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.

In fluid dynamics, aerodynamic potential flow codes or panel codes are used to determine the fluid velocity, and subsequently the pressure distribution, on an object. This may be a simple two-dimensional object, such as a circle or wing, or it may be a three-dimensional vehicle.

In fluid dynamics, the Oseen equations describe the flow of a viscous and incompressible fluid at small Reynolds numbers, as formulated by Carl Wilhelm Oseen in 1910. Oseen flow is an improved description of these flows, as compared to Stokes flow, with the (partial) inclusion of convective acceleration.

The convection–diffusion equation is a combination of the diffusion and convection (advection) equations, and describes physical phenomena where particles, energy, or other physical quantities are transferred inside a physical system due to two processes: diffusion and convection. Depending on context, the same equation can be called the advection–diffusion equation, drift–diffusion equation, or (generic) scalar transport equation.

<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, like in spinodal decomposition.

References

  1. Fliert, B. W. Van De; Howell, P. D.; Ockenden, J. R. (June 1995). "Pressure-driven flow of a thin viscous sheet". Journal of Fluid Mechanics. 292: 359–376. Bibcode:1995JFM...292..359V. doi:10.1017/S002211209500156X. ISSN   1469-7645. S2CID   120047555.
  2. Buckmaster, J. D.; Nachman, A.; Ting, L. (May 1975). "The buckling and stretching of a viscida". Journal of Fluid Mechanics. 69 (1): 1–20. Bibcode:1975JFM....69....1B. doi:10.1017/S0022112075001279. ISSN   1469-7645. S2CID   120390660.
  3. 1 2 A. Oron, S. H. Davis, S. G. Bankoff, "Long-scale evolution of thin liquid films", Rev. Mod. Phys., 69, 931–980 (1997)
  4. H. Knüpfer, "Classical solutions for a thin-film equation", PhD thesis, University of Bonn.
  5. 1 2 3 Myers, T. G. (January 1998). "Thin Films with High Surface Tension". SIAM Review. 40 (3): 441–462. Bibcode:1998SIAMR..40..441M. doi:10.1137/S003614459529284X. ISSN   0036-1445.
  6. O'Brien, S. B. G. M. (September 1993). "On Marangoni drying: nonlinear kinematic waves in a thin film". Journal of Fluid Mechanics. 254: 649–670. Bibcode:1993JFM...254..649O. doi:10.1017/S0022112093002290. ISSN   0022-1120. S2CID   122742594.
  7. Myers, T. G.; Charpin, J. P. F.; Thompson, C. P. (January 2002). "Slowly accreting ice due to supercooled water impacting on a cold surface". Physics of Fluids. 14 (1): 240–256. Bibcode:2002PhFl...14..240M. doi:10.1063/1.1416186. ISSN   1070-6631.
  8. Constantin, Peter; Dupont, Todd F.; Goldstein, Raymond E.; Kadanoff, Leo P.; Shelley, Michael J.; Zhou, Su-Min (1993-06-01). "Droplet breakup in a model of the Hele-Shaw cell". Physical Review E. 47 (6): 4169–4181. Bibcode:1993PhRvE..47.4169C. doi:10.1103/PhysRevE.47.4169. ISSN   1063-651X. PMID   9960494.
  9. 1 2 3 L. W. Schwartz, R. V. Roy, R. R. Eley, S. Petrash, "Dewetting patterns in a drying liquid film", Journal of Colloid and Interface Science, 243, 363374 (2001).
  10. Myers, T. G.; Charpin, J. P. F.; Chapman, S. J. (August 2002). "The flow and solidification of a thin fluid film on an arbitrary three-dimensional surface". Physics of Fluids. 14 (8): 2788–2803. Bibcode:2002PhFl...14.2788M. doi:10.1063/1.1488599. hdl: 2117/102903 . ISSN   1070-6631.
  11. Tuck, E. O.; Schwartz, L. W. (September 1990). "A Numerical and Asymptotic Study of Some Third-Order Ordinary Differential Equations Relevant to Draining and Coating Flows". SIAM Review. 32 (3): 453–469. doi:10.1137/1032079. ISSN   0036-1445.
  12. Balmforth, Neil; Ghadge, Shilpa; Myers, Tim (March 2007). "Surface tension driven fingering of a viscoplastic film". Journal of Non-Newtonian Fluid Mechanics. 142 (1–3): 143–149. doi:10.1016/j.jnnfm.2006.07.011.
  13. N.V. Churaev, V.D. Sobolev, Adv. Colloid Interface Sci. 61 (1995) 1-16
  14. 1 2 L. Kondic, "Instabilities in gravity driven flow of thin liquid films", SIAM Review, 45, 95–115 (2003)