Reynolds equation

Last updated

In fluid mechanics (specifically lubrication theory), the Reynolds equation is a partial differential equation governing the pressure distribution of thin viscous fluid films. It was first derived by Osborne Reynolds in 1886. [1] The classical Reynolds Equation can be used to describe the pressure distribution in nearly any type of fluid film bearing; a bearing type in which the bounding bodies are fully separated by a thin layer of liquid or gas.

Contents

General usage

The general Reynolds equation is:

Where:

The equation can either be used with consistent units or nondimensionalized.

The Reynolds Equation assumes:

For some simple bearing geometries and boundary conditions, the Reynolds equation can be solved analytically. Often however, the equation must be solved numerically. Frequently this involves discretizing the geometric domain, and then applying a finite technique - often FDM, FVM, or FEM.

Derivation from Navier-Stokes

A full derivation of the Reynolds Equation from the Navier-Stokes equation can be found in numerous lubrication text books. [2] [3]

Solution of Reynolds Equation

In general, Reynolds equation has to be solved using numerical methods such as finite difference, or finite element. In certain simplified cases, however, analytical or approximate solutions can be obtained. [4]

For the case of rigid sphere on flat geometry, steady-state case and half-Sommerfeld cavitation boundary condition, the 2-D Reynolds equation can be solved analytically. This solution was proposed by a Nobel Prize winner Pyotr Kapitsa. Half-Sommerfeld boundary condition was shown to be inaccurate and this solution has to be used with care.

In case of 1-D Reynolds equation several analytical or semi-analytical solutions are available. In 1916 Martin obtained a closed form solution [5] for a minimum film thickness and pressure for a rigid cylinder and plane geometry. This solution is not accurate for the cases when the elastic deformation of the surfaces contributes considerably to the film thickness. In 1949, Grubin obtained an approximate solution [6] for so called elasto-hydrodynamic lubrication (EHL) line contact problem, where he combined both elastic deformation and lubricant hydrodynamic flow. In this solution it was assumed that the pressure profile follows Hertz solution. The model is therefore accurate at high loads, when the hydrodynamic pressure tends to be close to the Hertz contact pressure. [7]

Applications

The Reynolds equation is used to model the pressure in many applications. For example:

Reynolds Equation adaptations - Average Flow Model

In 1978 Patir and Cheng introduced an average flow model, [8] [9] which modifies the Reynolds equation to consider the effects of surface roughness on lubricated contacts. The average flow model spans the regimes of lubrication where the surfaces are close together and/or touching. The average flow model applied "flow factors" to adjust how easy it is for the lubricant to flow in the direction of sliding or perpendicular to it. They also presented terms for adjusting the contact shear calculation. In these regimes, the surface topography acts to direct the lubricant flow, which has been demonstrated to affect the lubricant pressure and thus the surface separation and contact friction. [10]

Several notable attempts have been made to taken additional details of the contact into account in the simulation of fluid films in contacts. Leighton et al. [10] presented a method for determining the flow factors needed for the average flow model from any measured surface. Harp and Salent [11] extended the average flow model by considering the inter-asperity cavitation. Chengwei and Linqing [12] used an analysis of the surface height probability distribution to remove one of the more complex terms from the average Reynolds equation, and replace it with a flow factor referred to as contact flow factor, . Knoll et al. calculated flow factors, taking into account the elastic deformation of the surfaces. Meng et al. [13] also considered the elastic deformation of the contacting surfaces.

The work of Patir and Cheng was a precursor to the investigations of surface texturing in lubricated contacts. Demonstrating how large scale surface features generated micro-hydrodynamic lift to separate films and reduce friction, but only when the contact conditions support this. [14]

The average flow model of Patir and Cheng, [8] [9] is often coupled with the rough surface interaction model of Greenwood and Tripp [15] for modelling of the interaction of rough surfaces in loaded contacts. [10] [16]

Related Research Articles

<span class="mw-page-title-main">Fluid dynamics</span> Aspects of fluid mechanics involving flow

In physics, physical chemistry and engineering, fluid dynamics is a subdiscipline of fluid mechanics that describes the flow of fluids—liquids and gases. It has several subdisciplines, including aerodynamics and hydrodynamics. Fluid dynamics has a wide range of applications, including calculating forces and moments on aircraft, determining the mass flow rate of petroleum through pipelines, predicting weather patterns, understanding nebulae in interstellar space and modelling fission weapon detonation.

<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, named after French engineer and physicist Claude-Louis Navier and 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 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">Potential flow</span> Velocity field as the gradient of a scalar function

In fluid dynamics, potential flow describes the velocity field as the gradient of a scalar function: the velocity potential. As a result, a potential flow is characterized by an irrotational velocity field, which is a valid approximation for several applications. The irrotationality of a potential flow is due to the curl of the gradient of a scalar always being equal to zero.

<span class="mw-page-title-main">Boundary layer</span> Layer of fluid in the immediate vicinity of a bounding surface

In physics and fluid mechanics, a boundary layer is the thin layer of fluid in the immediate vicinity of a bounding surface formed by the fluid flowing along the surface. The fluid's interaction with the wall induces a no-slip boundary condition. The flow velocity then monotonically increases above the surface until it returns to the bulk flow velocity. The thin layer consisting of fluid whose velocity has not yet returned to the bulk flow velocity is called the velocity boundary layer.

The primitive equations are a set of nonlinear partial differential equations that are used to approximate global atmospheric flow and are used in most atmospheric models. They consist of three main sets of balance equations:

  1. A continuity equation: Representing the conservation of mass.
  2. Conservation of momentum: Consisting of a form of the Navier–Stokes equations that describe hydrodynamical flow on the surface of a sphere under the assumption that vertical motion is much smaller than horizontal motion (hydrostasis) and that the fluid layer depth is small compared to the radius of the sphere
  3. A thermal energy equation: Relating the overall temperature of the system to heat sources and sinks
<span class="mw-page-title-main">Lubrication</span> The presence of a material to reduce friction between two surfaces.

Lubrication is the process or technique of using a lubricant to reduce friction and wear and tear in a contact between two surfaces. The study of lubrication is a discipline in the field of tribology.

In fluid mechanics or more generally continuum mechanics, incompressible flow refers to a flow in which the material density is constant within a fluid parcel—an infinitesimal volume that moves with the flow velocity. An equivalent statement that implies incompressibility is that the divergence of the flow velocity is zero.

<span class="mw-page-title-main">D'Alembert's paradox</span>

In fluid dynamics, d'Alembert's paradox is a contradiction reached in 1752 by French mathematician Jean le Rond d'Alembert. D'Alembert proved that – for incompressible and inviscid potential flow – the drag force is zero on a body moving with constant velocity relative to the fluid. Zero drag is in direct contradiction to the observation of substantial drag on bodies moving relative to fluids, such as air and water; especially at high velocities corresponding with high Reynolds numbers. It is a particular example of the reversibility paradox.

The Reynolds-averaged Navier–Stokes equations are time-averaged equations of motion for fluid flow. The idea behind the equations is Reynolds decomposition, whereby an instantaneous quantity is decomposed into its time-averaged and fluctuating quantities, an idea first proposed by Osborne Reynolds. The RANS equations are primarily used to describe turbulent flows. These equations can be used with approximations based on knowledge of the properties of flow turbulence to give approximate time-averaged solutions to the Navier–Stokes equations. For a stationary flow of an incompressible Newtonian fluid, these equations can be written in Einstein notation in Cartesian coordinates as:

<span class="mw-page-title-main">Smoothed-particle hydrodynamics</span> Method of hydrodynamics simulation

Smoothed-particle hydrodynamics (SPH) is a computational method used for simulating the mechanics of continuum media, such as solid mechanics and fluid flows. It was developed by Gingold and Monaghan and Lucy in 1977, initially for astrophysical problems. It has been used in many fields of research, including astrophysics, ballistics, volcanology, and oceanography. It is a meshfree Lagrangian method, and the resolution of the method can easily be adjusted with respect to variables such as density.

<span class="mw-page-title-main">Rayleigh–Taylor instability</span> Unstable behavior of two contacting fluids of different densities

The Rayleigh–Taylor instability, or RT instability, is an instability of an interface between two fluids of different densities which occurs when the lighter fluid is pushing the heavier fluid. Examples include the behavior of water suspended above oil in the gravity of Earth, mushroom clouds like those from volcanic eruptions and atmospheric nuclear explosions, supernova explosions in which expanding core gas is accelerated into denser shell gas, instabilities in plasma fusion reactors and inertial confinement fusion.

In fluid dynamics, the Reynolds stress is the component of the total stress tensor in a fluid obtained from the averaging operation over the Navier–Stokes equations to account for turbulent fluctuations in fluid momentum.

<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">Shallow water equations</span> Set of partial differential equations that describe the flow below a pressure surface in a fluid

The shallow-water equations (SWE) are a set of hyperbolic partial differential equations that describe the flow below a pressure surface in a fluid. The shallow-water equations in unidirectional form are also called Saint-Venant equations, after Adhémar Jean Claude Barré de Saint-Venant.

The Stribeck curve is a fundamental concept in the field of tribology. It shows that friction in fluid-lubricated contacts is a non-linear function of the contact load, the lubricant viscosity and the lubricant entrainment speed. The discovery and underlying research is usually attributed to Richard Stribeck and Mayo D. Hersey, who studied friction in journal bearings for railway wagon applications during the first half of the 20th century; however, other researchers have arrived at similar conclusions before. The mechanisms along the Stribeck curve have been understood today also on the atomistic level.

In fluid dynamics, Airy wave theory gives a linearised description of the propagation of gravity waves on the surface of a homogeneous fluid layer. The theory assumes that the fluid layer has a uniform mean depth, and that the fluid flow is inviscid, incompressible and irrotational. This theory was first published, in correct form, by George Biddell Airy in the 19th century.

<span class="mw-page-title-main">Hydrodynamic stability</span> Subfield of fluid dynamics

In fluid dynamics, hydrodynamic stability is the field which analyses the stability and the onset of instability of fluid flows. The study of hydrodynamic stability aims to find out if a given flow is stable or unstable, and if so, how these instabilities will cause the development of turbulence. The foundations of hydrodynamic stability, both theoretical and experimental, were laid most notably by Helmholtz, Kelvin, Rayleigh and Reynolds during the nineteenth century. These foundations have given many useful tools to study hydrodynamic stability. These include Reynolds number, the Euler equations, and the Navier–Stokes equations. When studying flow stability it is useful to understand more simplistic systems, e.g. incompressible and inviscid fluids which can then be developed further onto more complex flows. Since the 1980s, more computational methods are being used to model and analyse the more complex flows.

<span class="mw-page-title-main">Reynolds number</span> Ratio of inertial to viscous forces acting on a liquid

In fluid mechanics, the Reynolds number is a dimensionless quantity that helps predict fluid flow patterns in different situations by measuring the ratio between inertial and viscous forces. At low Reynolds numbers, flows tend to be dominated by laminar (sheet-like) flow, while at high Reynolds numbers, flows tend to be turbulent. The turbulence results from differences in the fluid's speed and direction, which may sometimes intersect or even move counter to the overall direction of the flow. These eddy currents begin to churn the flow, using up energy in the process, which for liquids increases the chances of cavitation.

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.

References

  1. Reynolds, O. (1886). "On the Theory of Lubrication and Its Application to Mr. Beauchamp Tower's Experiments, Including an Experimental Determination of the Viscosity of Olive Oil". Philosophical Transactions of the Royal Society of London. Royal Society. 177: 157–234. doi:10.1098/rstl.1886.0005. JSTOR   109480. S2CID   110829869.
  2. Hamrock, Bernard J.; Schmid, Steven R.; Jacobson, Bo O. (2004). Fundamentals of Fluid Film Lubrication. Taylor & Francis. ISBN   978-0-8247-5371-9.
  3. Szeri, Andras Z. (2010). Fluid Film Lubrication. Cambridge University Press. ISBN   978-0-521-89823-2.
  4. "Reynolds Equation: Derivation and Solution". tribonet.org. 12 November 2016. Retrieved 10 September 2019.
  5. Akchurin, Aydar (18 February 2016). "Analytical Solution of 1D Reynolds Equation". tribonet.org. Retrieved 10 September 2019.
  6. Akchurin, Aydar (22 February 2016). "Semi-Analytical Solution of 1D Transient Reynolds Equation(Grubin's Approximation)". tribonet.org. Retrieved 10 September 2019.
  7. Akchurin, Aydar (4 January 2017). "Hertz Contact Calculator". tribonet.org. Retrieved 10 September 2019.
  8. 1 2 Patir, Nadir; Cheng, H. S. (1978). "An Average Flow Model for Determining Effects of Three-Dimensional Roughness on Partial Hydrodynamic Lubrication". Journal of Lubrication Technology. 100 (1): 12. doi:10.1115/1.3453103. ISSN   0022-2305.
  9. 1 2 Patir, Nadir; Cheng, H. S. (1979-04-01). "Application of Average Flow Model to Lubrication Between Rough Sliding Surfaces". Journal of Lubrication Technology. 101 (2): 220–229. doi:10.1115/1.3453329. ISSN   0022-2305.
  10. 1 2 3 Leighton; et al. (2016). "Surface-specific flow factors for prediction of friction of cross-hatched surfaces". Surface Topography: Metrology and Properties. 4 (2): 025002. doi: 10.1088/2051-672x/4/2/025002 . S2CID   111631084.
  11. Harp, Susan R.; Salant, Richard F. (2000-10-17). "An Average Flow Model of Rough Surface Lubrication With Inter-Asperity Cavitation". Journal of Tribology. 123 (1): 134–143. doi:10.1115/1.1332397. ISSN   0742-4787.
  12. Wu, Chengwei; Zheng, Linqing (1989-01-01). "An Average Reynolds Equation for Partial Film Lubrication With a Contact Factor". Journal of Tribology. 111 (1): 188–191. doi:10.1115/1.3261872. ISSN   0742-4787.
  13. Meng, F-M; Wang, W-Z; Hu, Y-Z; Wang, H (2007-07-01). "Numerical analysis of combined influences of inter-asperity cavitation and elastic deformation on flow factors". Proceedings of the Institution of Mechanical Engineers, Part C: Journal of Mechanical Engineering Science. 221 (7): 815–827. doi:10.1243/0954406jmes525. ISSN   0954-4062. S2CID   137022386.
  14. Morris, N; Leighton, M; De la Cruz, M; Rahmani, R; Rahnejat, H; Howell-Smith, S (2014-11-17). "Combined numerical and experimental investigation of the micro-hydrodynamics of chevron-based textured patterns influencing conjunctional friction of sliding contacts". Proceedings of the Institution of Mechanical Engineers, Part J: Journal of Engineering Tribology. 229 (4): 316–335. doi: 10.1177/1350650114559996 . ISSN   1350-6501. S2CID   53586245.
  15. Greenwood, J. A.; Tripp, J. H. (June 1970). "The Contact of Two Nominally Flat Rough Surfaces". Proceedings of the Institution of Mechanical Engineers. 185 (1): 625–633. doi:10.1243/pime_proc_1970_185_069_02. ISSN   0020-3483.
  16. Leighton, M; Nicholls, T; De la Cruz, M; Rahmani, R; Rahnejat, H (2016-12-12). "Combined lubricant–surface system perspective: Multi-scale numerical–experimental investigation". Proceedings of the Institution of Mechanical Engineers, Part J: Journal of Engineering Tribology. 231 (7): 910–924. doi: 10.1177/1350650116683784 . ISSN   1350-6501. S2CID   55438508.