Fact-checked by Grok 2 weeks ago

Lubrication theory

Lubrication theory is a fundamental framework in that governs the flow of viscous lubricants in thin films separating two surfaces in relative motion, enabling the generation of hydrodynamic to support loads while minimizing and through complete surface separation. This theory simplifies the Navier-Stokes equations under assumptions of low , negligible inertia, and small gap thickness relative to lateral dimensions, yielding the as its cornerstone for predicting distribution and film thickness. Formally established by Osborne Reynolds in 1886 through analysis of journal bearing experiments, it explains how viscous fluids like oils reduce direct solid contact in mechanical systems. The theory encompasses several lubrication regimes, distinguished by the film parameter \Lambda = h_{\min} / \sqrt{R_{q,a}^2 + R_{q,b}^2}, where h_{\min} is the minimum film thickness and R_q represents surface roughness. In hydrodynamic lubrication (\Lambda \geq 5), rigid surfaces are fully separated by a lubricant film generated solely by motion, with no elastic deformation considered; applications include journal and thrust bearings. Elastohydrodynamic lubrication (EHL) ($3 \leq \Lambda < 10) extends this to nonconformal contacts like gears and rolling-element bearings, incorporating elastic deformation of surfaces and pressure-dependent lubricant viscosity, often using empirical formulas such as Hamrock and Dowson's for minimum film thickness: H_{\min} = 3.63 U^{0.68} G^{0.49} W^{-0.073} (1 - e^{-0.68k}), where U, G, and W are dimensionless speed, material, and load parameters, respectively. Boundary lubrication (\Lambda < 1) involves partial film breakdown with asperity contact mitigated by boundary films, though it receives less emphasis in core theory. Historically, precursors include Newton's 1687 viscosity postulate and Tower's 1885 experiments demonstrating pressure generation in lubricated bearings, but Reynolds' derivation unified these into a predictive model applicable to diverse engineering contexts, from automotive transmissions to biological joints. Modern extensions address non-Newtonian effects, thermal influences, and compressible gases, enhancing design for energy efficiency and durability in high-speed machinery. By ensuring adequate film thickness—proportional to speed and inversely to load—lubrication theory prevents fatigue and extends component life, as validated in endurance tests of ball bearings.

Introduction

Definition and Scope

Lubrication theory is a branch of fluid mechanics dedicated to the analysis of viscous flows in thin-film geometries, where the characteristic film thickness H is significantly smaller than the length L along the flow direction, resulting in a small aspect ratio \epsilon = H/L \ll 1. This approximation simplifies the governing equations to model pressure-driven lubrication flows between closely spaced surfaces, such as those in mechanical contacts. The theory originated from efforts to understand how viscous fluids maintain separation and reduce friction in such configurations. The scope of lubrication theory extends to both liquid and gaseous lubricants, with a primary emphasis on hydrodynamic lubrication, where a complete fluid film fully separates opposing surfaces under motion, preventing direct solid-solid contact. It contrasts with boundary lubrication regimes, in which partial surface asperity contact occurs due to insufficient film thickness. The core objectives include predicting the pressure distribution across the lubricant film, assessing load-carrying capacity, and evaluating friction minimization to ensure efficient operation. These predictions arise from asymptotic simplifications of the , often leading to the as the central governing relation. Typical geometries analyzed include parallel plates, journal bearings, and slider bearings, where the thin-film assumption holds effectively. The theory's applicability spans scales, from microscale phenomena in nanofluidics and MEMS devices to macroscale systems like turbomachinery and heavy industrial bearings. Lubrication theory plays a critical role in engineering by enabling the design of systems that minimize wear, extend component life, and reduce energy dissipation through optimized fluid films. It serves as a foundational pillar of , the broader science of interacting surfaces in relative motion, influencing advancements in mechanical reliability across industries.

Historical Development

Practical applications of lubrication date back to ancient civilizations, with evidence from around 3500 BCE showing Egyptians and Sumerians employing animal fats, vegetable oils, and water to reduce friction on wheeled carts and sleds for transporting heavy loads such as building materials. By 2600 BCE, analysis of a sled wheel from an Egyptian pharaoh's tomb revealed the use of beef or ram tallow as a lubricant, demonstrating early recognition of viscous substances to ease movement over surfaces. These empirical uses persisted for millennia without a formal theoretical framework, as lubrication was treated primarily as a practical engineering solution rather than a scientific discipline. The theoretical foundations of lubrication emerged in the 19th century amid the Industrial Revolution's demand for efficient machinery. In 1883, Beauchamp Tower conducted pivotal experiments on journal bearings for the British railway system, observing that hydrodynamic pressure generated within the lubricant film supported the load and prevented direct metal-to-metal contact, thus explaining reduced friction without external pressurization. Building on Tower's findings, Osborne Reynolds published his seminal 1886 paper, "On the Theory of Lubrication and Its Application to Mr. Beauchamp Tower’s Experiments," deriving the fundamental equations governing thin-film lubrication from the Navier-Stokes equations under simplifying assumptions, establishing hydrodynamic lubrication as a rigorous field. Early 20th-century advancements refined Reynolds' theory through analytical solutions and empirical correlations. In 1904, Arnold Sommerfeld solved the Reynolds equation for infinitely long journal bearings, introducing key boundary conditions and providing closed-form expressions for pressure distribution and load capacity that became foundational for bearing design. Around the same period, Richard Stribeck's research at the Technical University of Berlin in the early 1900s, which influenced bearing designs at companies like SKF, led to the Stribeck curve, which experimentally linked friction coefficient to lubrication regimes—boundary, mixed, and hydrodynamic—based on viscosity, speed, and load, bridging theory with practical friction behavior. Mid-20th-century developments addressed limitations in classical hydrodynamic theory, incorporating elasticity and thermal effects. In the 1940s, Alexander Ertel and A. Grubin introduced elastohydrodynamic lubrication (EHL) theory to explain film formation in high-pressure, non-conformal contacts like gears and rollers, where elastic deformation significantly influences film thickness. Analytical advancements, such as the 1949 work by Milton C. Shaw and E. Fred Macks on bearing lubrication, further supported theoretical predictions on load-carrying capacity and film stability. By the 1950s, studies on thermal effects highlighted the role of lubricant heating in reducing viscosity and altering performance, prompting extensions to the Reynolds equation for non-isothermal conditions. In the 1960s, Jakob A. Greenwood advanced the framework by formalizing three-dimensional solutions to the Reynolds equation for finite bearings and rough surfaces, enabling more accurate modeling of real-world geometries and contact mechanics. In the late 1960s, Dowson and Higginson developed numerical solutions and empirical formulas for EHL film thickness, building on Grubin's approximations.

Fundamental Principles

Key Assumptions and Approximations

Lubrication theory relies on the core approximation that the fluid film is thin compared to its lateral extent, characterized by the aspect ratio \epsilon = H/L \ll 1, where H is the characteristic film thickness and L is the streamwise length scale. This small \epsilon implies that viscous forces dominate over inertial forces, allowing the neglect of inertia terms in the governing equations, and that pressure gradients are primarily in the streamwise direction (x) while transverse variations (\partial p / \partial z \approx 0) are negligible across the film thickness. The theory assumes a Newtonian, incompressible fluid with constant viscosity, adhering to no-slip boundary conditions at the solid surfaces. Flow is steady-state and isothermal, with no significant heat generation, and body forces are negligible except for gravity in free-surface films. Surface tension effects are typically ignored but become relevant in free-surface flows for films thinner than approximately 1 \mum. These assumptions stem from Osborne Reynolds' foundational 1886 analysis, which simplified the for thin viscous films between bearing surfaces. Scaling analysis underpins these approximations. The streamwise velocity scales as u \sim U, where U is the characteristic surface speed, while the transverse velocity scales as w \sim \epsilon U. The pressure scales as p \sim \mu U L / H^2, where \mu is the dynamic viscosity, leading to viscous stresses \tau \sim \mu U / H that balance the pressure gradient in the streamwise momentum equation. A reduced Reynolds number \mathrm{Re}^* = \rho U H / \mu \ll 1/\epsilon ensures inertial terms are O(\epsilon^2) smaller than viscous terms, justifying their omission, and the transverse momentum balance confirms \partial p / \partial z \approx 0. Typically, \epsilon \sim 10^{-3}, making these scalings valid for many engineering films. The approximations break down at high speeds where inertial effects become significant (\mathrm{Re}^* \gtrsim 1/\epsilon), or in extremely thin films where molecular or continuum assumptions fail, such as when the film thickness approaches the mean free path for gases or a few molecular layers for liquids. In such cases, full computational fluid dynamics or modified theories are required instead of the lubrication approximation.

Internal vs. Free-Surface Flows

In lubrication theory, internal flows refer to scenarios where the lubricating fluid is fully confined between two solid surfaces, which may be rigid or deformable, such as in or . The primary objective in analyzing these flows is to determine the pressure distribution p(x,y) across the fluid film, which enables the computation of the load-carrying capacity W = \int p \, dA and the attitude angle that describes the orientation of the load relative to the bearing surfaces. Unlike other flow configurations, internal flows lack free boundaries, allowing the focus to remain on the force balance between the solids mediated by the viscous fluid without additional interface dynamics. In contrast, free-surface lubrication flows involve a fluid domain bounded by a solid substrate on one side and a free interface on the other, as seen in processes like or the flow of a viscous liquid down an inclined plane. These flows require solving for the evolution of the interface height h(x,t), governed by the kinematic condition \frac{\partial h}{\partial t} + u \frac{\partial h}{\partial x} = w, where u and w are the horizontal and vertical velocity components at the interface, respectively. Additionally, surface tension effects are incorporated through the curvature-driven term \nabla \cdot (\sigma \mathbf{n}), where \sigma is the surface tension and \mathbf{n} is the interface normal, while wetting dynamics are captured by the contact angle \theta at the three-phase contact line. The key differences between internal and free-surface flows arise from their distinct physical emphases: internal flows prioritize the equilibrium of forces on the confining solids to ensure stable separation and load support, whereas free-surface flows introduce time-dependent evolution equations for the interface height h, potentially leading to instabilities like dewetting in ultra-thin films driven by Van der Waals forces. For instance, a viscous gravity current spreading over a horizontal surface exemplifies free-surface dynamics, where buoyancy and viscosity dictate the front propagation, in contrast to a sealed journal bearing, which maintains a closed fluid domain focused on pressure-induced hydrodynamic support. In free-surface flows, intermolecular effects become prominent when the film thickness h falls below 1 \mum, introducing a disjoining pressure \Pi(h) primarily from Van der Waals interactions that can modify the effective viscosity or induce film instability and rupture. This disjoining pressure, often modeled as \Pi(h) = -\frac{A}{6\pi h^3} where A is the Hamaker constant, alters the pressure gradient in the lubrication equation, promoting either stabilization or dewetting depending on the sign and magnitude of A.

Mathematical Formulation

Derivation from Navier-Stokes Equations

Lubrication theory begins with the incompressible for a Newtonian fluid, which describe the conservation of momentum and mass: \rho \left( \frac{\partial \mathbf{u}}{\partial t} + \mathbf{u} \cdot \nabla \mathbf{u} \right) = -\nabla p + \mu \nabla^2 \mathbf{u} + \rho \mathbf{g}, \nabla \cdot \mathbf{u} = 0, where \mathbf{u} = (u, v, w) is the velocity vector, p is pressure, \rho is density, \mu is dynamic viscosity, and \mathbf{g} is the body force per unit mass (often gravity). These equations are simplified under the thin-film limit characteristic of lubrication flows, where the film thickness H is much smaller than the lateral dimensions L, defining the small aspect ratio \epsilon = H/L \ll 1. To derive the lubrication approximations, the equations are non-dimensionalized using characteristic scales appropriate to the geometry: x, y \sim L, z \sim H, u, v \sim U (lateral velocity scale), w \sim \epsilon U (vertical velocity scale), p \sim \mu U L / H^2 (viscous pressure scale), and t \sim L/U (convective time scale). Substituting these into the Navier-Stokes equations reveals the relative magnitudes of terms. The inertial terms scale as \mathrm{Re} \epsilon^2, where \mathrm{Re} = \rho U L / \mu is the Reynolds number; for typical lubrication regimes, \mathrm{Re} \epsilon^2 \ll 1, allowing neglect of inertia. The pressure gradient in the z-direction scales as \epsilon^2 times the lateral gradient, so \partial p / \partial z \sim \epsilon^2 \partial p / \partial x \ll \partial p / \partial x. Body forces are often negligible unless buoyancy or other effects dominate. The simplified momentum equations follow from this scaling. The z-momentum equation reduces to \partial p / \partial z = 0, implying pressure is constant across the film thickness: p = p(x, y). The x-momentum equation balances pressure gradient and viscous diffusion, yielding $0 = -\partial p / \partial x + \mu \partial^2 u / \partial z^2 (Poiseuille-Couette flow). The y-momentum equation is analogous: $0 = -\partial p / \partial y + \mu \partial^2 v / \partial z^2. Integrating the x-momentum equation twice with no-slip boundary conditions—at z = 0, u = 0 (stationary lower surface), and at z = h(x, y), u = U (upper surface moving in the x-direction, as in a simple slider bearing)—gives the velocity profile: u(z) = \frac{1}{2\mu} \frac{\partial p}{\partial x} (z^2 - z h) + U \frac{z}{h}. A similar expression holds for v(z). These profiles capture the parabolic pressure-driven (Poiseuille) component superimposed on the linear shear-driven (Couette) component. Integrating the continuity equation across the film thickness from z = 0 to z = h enforces mass conservation. For steady, two-dimensional flow (neglecting y-variation for simplicity), this yields \partial / \partial x \int_0^h u \, dz = 0, or more generally, the divergence of the depth-averaged flux \mathbf{q} = \int_0^h \mathbf{u} \, dz = 0. The x-component of the flux is q_x = \int_0^h u \, dz = -\frac{h^3}{12\mu} \frac{\partial p}{\partial x} + \frac{U h}{2}, with an analogous q_y. The first term represents diffusive transport due to pressure gradients, while the second is advective transport due to surface motion. For unsteady or three-dimensional cases, additional terms arise from film thickness changes, but the steady form \nabla \cdot \mathbf{q} = 0 provides the foundation for pressure determination in lubrication problems.

Reynolds Equation and Boundary Conditions

The governs the pressure distribution in thin lubricant films under the assumptions of , integrating the simplified across the film thickness. For steady, incompressible, internal flows between two surfaces moving with velocities U in the x-direction and V in the y-direction, the two-dimensional form is \frac{\partial}{\partial x} \left( \frac{h^3}{12\mu} \frac{\partial p}{\partial x} \right) + \frac{\partial}{\partial y} \left( \frac{h^3}{12\mu} \frac{\partial p}{\partial y} \right) = \frac{1}{2} \frac{\partial (U h)}{\partial x} + \frac{1}{2} \frac{\partial (V h)}{\partial y}, where h(x,y) is the film thickness, \mu is the constant dynamic viscosity, and p(x,y) is the pressure. This equation, first derived by Osborne Reynolds, balances the Poiseuille pressure-driven flow with the Couette shear-driven flow. For unsteady flows, a time-dependent term is added to the right-hand side: \frac{\partial h}{\partial t}, yielding the general three-dimensional form when extended to the z-direction if needed. In free-surface lubrication, such as coating flows or meniscus-driven films, the Reynolds equation couples with the kinematic boundary condition at the liquid-air interface to enforce mass conservation: \frac{\partial h}{\partial t} + \nabla \cdot \mathbf{q} = 0, where the depth-averaged flux \mathbf{q} = -\frac{h^3}{12\mu} \nabla p + \frac{U h}{2} incorporates both pressure gradients and mean surface velocity U. This formulation applies to thin films where surface tension effects dominate at the edges. Boundary conditions are essential for solving the Reynolds equation and reflect physical constraints at the film domain. For internal flows, ambient pressure p = p_a (typically p_a = 0 gauge) is imposed at the inlet and outlet edges where the film opens to the surroundings. In regions prone to cavitation, such as journal bearings under load, the pressure cannot drop below the vapor pressure (often approximated as p \geq 0); the Jakobsson-Floberg-Olsson (JFO) model handles this by switching off the diffusive terms in cavitated zones, enforcing p = 0 and \frac{\partial p}{\partial n} = 0 (normal derivative zero) across the film-rupture interface while conserving mass flux. For free surfaces, the dynamic boundary condition arises from the stress balance \boldsymbol{\sigma} \cdot \mathbf{n} = -p \mathbf{n} + \nabla \cdot (\sigma \nabla h), where \boldsymbol{\sigma} is the stress tensor, \mathbf{n} the normal, and \sigma the surface tension; this simplifies to the Young-Laplace equation p = \sigma \kappa at the interface, with \kappa the mean curvature. Solutions to the Reynolds equation range from analytical for idealized cases to numerical for practical geometries. In simple one-dimensional slider bearings with constant viscosity and no side leakage, an analytical pressure profile is p(x) = \frac{6\mu U L^2}{h^2} \left( \frac{x}{L} \right) \left(1 - \frac{x}{L} \right), illustrating the parabolic pressure build-up that supports load via the . Complex configurations, including variable film thickness, , or irregular boundaries, require numerical methods such as , , or to discretize and solve the elliptic-parabolic iteratively.

Applications

Engineering and Mechanical Systems

Lubrication theory plays a pivotal role in the design and operation of hydrodynamic bearings, which support rotating shafts in mechanical systems by generating a pressurized lubricant film to separate surfaces and carry loads. Journal bearings, consisting of a cylindrical shaft rotating within a sleeve, rely on the wedge-shaped film formed by shaft motion and eccentricity to develop hydrodynamic pressure. The load-carrying capacity W from solutions to the Reynolds equation under typical operating conditions is W = \frac{\mu U r L}{c^2} f(\epsilon), where \mu is the lubricant viscosity, U is the sliding speed, r is the journal radius, L is the bearing length, c is the radial clearance, \epsilon is the eccentricity ratio, and f(\epsilon) is a dimensionless function depending on bearing geometry (long or short approximation). Thrust bearings, which handle axial loads in applications like turbines, employ tilted pads or sectors to create converging films, achieving similar load support through hydrodynamic action while accommodating misalignment. The friction coefficient f in hydrodynamic journal bearings is approximately given by Petroff's law for lightly loaded conditions: f \approx 2\pi \left( \frac{\mu U}{W} \right) \left( \frac{r}{c} \right), or more generally from solutions involving the Sommerfeld number. In mechanical seals and gears, lubrication theory ensures leakage prevention by maintaining thin hydrodynamic films that seal interfaces under pressure differentials. Mechanical face seals generate a fluid film via relative motion to minimize leakage, with film thickness controlling flow rates proportional to pressure and inversely to viscosity. For gears, thin films separate meshing teeth, reducing wear and containing lubricants; the Stribeck curve delineates the transition from boundary lubrication (high friction, direct contact) to hydrodynamic regimes (low friction, full separation) as a function of the parameter \eta N / P (lubricant viscosity \eta, speed N, load P), guiding selection of operating conditions to avoid mixed regimes where partial contact occurs. In rotating machinery such as turbines and automotive engines, lubrication theory predicts stability against phenomena like rotor whirl, where hydrodynamic forces can destabilize the system at supercritical speeds, leading to subsynchronous vibrations if film stiffness and damping are insufficient. Theory aids in ensuring minimum film thickness h_{\min} > 1 \, \mu \mathrm{m} to prevent metal-to-metal contact and , particularly under transient loads in engines where speeds exceed 10,000 rpm. implications emphasize optimizing the eccentricity ratio \varepsilon_b = e/c, typically targeting values around 0.6–0.8 for balanced load capacity and , as higher eccentricity enhances peaks but risks film rupture. Historical validation through Beauchamp Tower's 1883–1884 experiments demonstrated generation in bearings under load, confirming hydrodynamic principles and inspiring Reynolds' foundational work.

Biological and Coating Processes

Lubrication theory extends to biological systems where deformable, permeable surfaces and thin layers interact under low conditions, often involving free surfaces or evolving interfaces. In synovial joints, articular acts as a permeable bearing that facilitates low-friction motion through biphasic lubrication mechanisms, including squeeze- action where interstitial pressurization supports loads and reduces contact stresses. The effective thickness in these joints can approach 100 nm in boundary regimes, enabling near-frictionless sliding via molecular interactions at the surface. This permeable nature allows to filter through the matrix, enhancing load-bearing capacity during compression, as modeled by asymptotic analyses of squeeze- dynamics under physiological loading. Blood flow in capillaries exemplifies lubrication principles through the Fåhraeus-Lindqvist effect, where the apparent viscosity of blood decreases in narrow vessels (diameters below 300 μm) due to a cell-free plasma layer near the wall, reducing effective shear resistance. This results in an effective viscosity η_eff lower than the bulk plasma viscosity η, optimizing flow in microvascular networks by minimizing energy dissipation. The effect arises from red blood cell migration away from walls, creating a low-viscosity marginal zone that aligns with lubrication approximations for thin-film flows. In the lungs, films at the alveolar air-liquid interface apply lubrication theory to stabilize thin films against , reducing from ~72 mN/m to near-zero during compression to prevent . These phospholipid-based monolayers, secreted by type II alveolar cells, form viscoelastic layers that resist film rupture under cyclic , modeled as free-surface flows where Marangoni stresses balance gravitational and forces. This dynamic modulation ensures alveolar stability, with deficiencies leading to respiratory distress in neonates. Coating processes leverage lubrication theory for controlled deposition of thin in industrial applications, particularly and coating where viscous dominates. In coating, a substrate withdrawn from a at speed U entrains a whose thickness h scales as h \approx 0.64 R \left( \frac{\mu U}{\sigma} \right)^{2/3}, where R is the cylinder radius for thin fibers. coating similarly predicts uniform thicknesses at low (<10^{-2}), where from the meters the , achieving h \sim \mathrm{Ca}^{2/3} l_c , where l_c = \sqrt{\sigma / (\rho g)} is the . In and deposition, these models uniform layer formation, ensuring consistent wetting and minimizing defects in roll-to-roll processes. Free-surface dynamics in lubrication theory govern the of thin films under or , distinct from rigid-boundary cases. For -driven films on inclined planes, the Nusselt solution yields a uniform flow speed U = (g sin θ h^2)/(3 ν), where ν is kinematic , describing steady drainage balanced by and viscous shear. Dewetting instabilities in thin films arise from van der Waals forces destabilizing uniform layers below ~100 nm, leading to spinodal rupture or hole growth at rates predicted by equations, with rim formation and coarsening dynamics following power-law scaling in time. These processes highlight the role of intermolecular potentials in driving interface . Unique to biological contexts are deformable substrates and nanoscale forces that modify classical . In synovial joints, behaves as a poroelastic medium, where exudation from the biphasic couples with thin-film to sustain pressurization and low , contrasting rigid bearings. At nano-scales in bio-, intermolecular forces such as shells around glycoproteins enable boundary with coefficients below 0.001, where repulsive potentials prevent direct asperity contact in -on- interfaces.

Advanced Developments

Thermal and Compressibility Effects

In lubrication theory, thermal effects become significant in high-speed or heavily loaded systems where viscous dissipation generates substantial within the , leading to variations that alter properties. The primary mechanism is viscous heating, captured by the energy equation in the thin- : \rho c_p (\mathbf{u} \cdot \nabla T) = k \nabla^2 T + \mu \left( \frac{\partial u}{\partial z} \right)^2, where \rho is , c_p is specific , \mathbf{u} is , T is , k is thermal conductivity, \mu is , and the last term represents viscous dissipation dominant across the film thickness z. This couples to the momentum equations through temperature-dependent , often modeled as \mu(T) = \mu_0 \exp(-\beta (T - T_0)), where \mu_0 is reference , \beta is the temperature-viscosity (typically 0.02–0.05 K^{-1} for mineral oils), and T_0 is reference . The resulting reduction softens the pressure generation, decreasing load capacity; in high-speed bearings, this can cause a significant drop compared to isothermal predictions. For gaseous lubricants, compressibility introduces density variations that must be accounted for, extending the incompressible Reynolds equation. The continuity equation becomes \nabla \cdot (\rho \mathbf{u}) = 0, with density \rho related to pressure p via the isentropic assumption p / \rho^\gamma = \ constant, where \gamma is the specific heat ratio (1.4 for air). Integrating across the film yields the compressible Reynolds equation: \frac{\partial}{\partial x} \left( \rho h^3 \frac{\partial p}{\partial x} \right) = 12 \mu \frac{\partial (\rho U h)}{\partial x}, for one-dimensional sliding with surface velocity U and film thickness h, assuming isothermal conditions initially. This form highlights how density gradients enhance pressure buildup in converging films but require low Mach numbers, Ma = U / \sqrt{\gamma R T} < 1 (with gas constant R), to avoid shock waves in high-speed applications like air bearings. Coupled thermo-hydrodynamic models address both effects simultaneously through iterative solutions of the modified Reynolds, energy, and state equations, often numerically for accuracy. These are essential for high-speed gas-lubricated systems, such as air-lubricated journal bearings, where thermal-viscous coupling can limit performance; historical developments in the 1950s by W. A. Gross established foundational gas bearing theory, incorporating compressibility for stable operation under partial arc loading. In practice, such models predict reduced stiffness and damping when Ma approaches unity, guiding designs in turbines and precision machinery. Limitations of these extensions include neglect of turbulence at high Reynolds numbers and the need for two-dimensional numerical integration to capture circumferential variations, as one-dimensional approximations overestimate load capacity by ignoring heat conduction to boundaries.

Non-Newtonian Fluids and Elastohydrodynamics

Lubrication theory traditionally assumes Newtonian fluids with constant viscosity, but many practical lubricants exhibit non-Newtonian behavior, where viscosity depends on shear rate, necessitating modifications to the governing equations. For power-law fluids, the shear stress is modeled as \tau = K \left( \frac{\partial u}{\partial z} \right)^n, where K is the consistency index and n is the power-law exponent; values of n < 1 characterize shear-thinning greases, which reduce effective viscosity under high shear. The Carreau model provides a more comprehensive description for lubricants transitioning between zero-shear and infinite-shear viscosities: \eta(\dot{\gamma}) = \eta_\infty + (\eta_0 - \eta_\infty) [1 + (\lambda \dot{\gamma})^2]^{(n-1)/2}, capturing both shear-thinning and Newtonian plateaus relevant to journal bearings under varying speeds. To accommodate these rheologies, the Reynolds equation is generalized; for power-law fluids in one dimension, the flow rate becomes q = -\frac{n}{2n+1} \left( \frac{h^{2n+1}}{K} \right)^{1/n} \left| \frac{\partial p}{\partial x} \right|^{(1-n)/n} \frac{\partial p}{\partial x}, derived by integrating the momentum equation across the film thickness, which alters pressure distribution and load capacity in non-Newtonian flows. Shear-thinning effects (n < 1) in greases can lead to thinner films compared to Newtonian fluids under certain conditions by reducing effective viscosity, affecting performance in lubricated contacts. Elastohydrodynamic lubrication (EHL) extends classical theory by coupling hydrodynamic to elastic deformation of contacting surfaces, critical for high-load scenarios where rigid-body assumptions fail. The elastic displacement is given by the Hertzian integral u_e = \frac{2}{\pi E'} \int \frac{p(\xi)}{|x - \xi|} d\xi, where E' is the effective , leading to modified film thickness h = h_0 + u_e. The central scales as p_0 \sim \sqrt{\frac{W E'}{R^2}}, with R as the equivalent radius and W the load, reflecting the Hertzian contact . In EHL, the piezoviscous effect further complicates the flow, with increasing exponentially under pressure: \eta(p) = \eta_0 \exp(\alpha p), where \alpha is the pressure-viscosity coefficient, which sustains thin films (nanometers) in high-pressure zones by resisting . This is pivotal for applications like rolling element bearings and operating at high speeds and loads, where EHL prevents direct asperity contact and reduces . A seminal milestone is the Dowson-Higginson formula for central film thickness in point contacts: h_c = 2.69 (U \eta_0)^{0.67} (\alpha)^{0.53} (W')^{-0.067} (E')^{-0.03} R^{0.43}, empirically derived from numerical solutions and widely used for design predictions. Numerical advances in EHL leverage multigrid solvers to handle the coupled of Reynolds and elasticity equations, achieving efficient for point and line contacts by coarsening grids and accelerating iterations, as demonstrated in simulations of smooth and rough surfaces. Modern extensions include 21st-century bio-EHL, where natural biomacromolecules like components are modeled for joint lubrication, addressing gaps in classical theory for biological systems with viscoelastic surfaces.

References

  1. [1]
    [PDF] Fundamentals of Fluid Film Lubrication
    Fluid film lubrication occurs when opposing bearing surfaces are separated by a lubricant film, including hydrodynamic and elastohydrodynamic lubrication.
  2. [2]
    IV. On the theory of lubrication and its application to Mr. Beauchamp ...
    Lubrication is the action of oils and other viscous fluids to diminish friction and wear between solid surfaces.
  3. [3]
    Dynamics and stability of thin liquid films | Rev. Mod. Phys.
    Aug 5, 2009 · The key parameter ϵ ≡ H ∕ L , the film aspect ratio, is often referred to as the “lubrication parameter” and is taken asymptotically small ...
  4. [4]
    Hydrodynamic Lubrication - an overview | ScienceDirect Topics
    Hydrodynamic lubrication (HL) is defined as a lubrication method where a thick fluid film prevents direct contact between friction-coupled elements, enabling ...
  5. [5]
    A cross-scale analysis method for lubrication characteristics of micro ...
    Oct 13, 2025 · Micro-textured water-lubricated thrust bearings exhibit significant potential in motorized spindles due to their low friction, ...
  6. [6]
    Hydrodynamic Lubrication (HL): Theory and Definition - Tribonet
    The theory is developed within a field known as tribology. Lubricant is a substance which is used to control (more often to reduce) friction and wear of the ...
  7. [7]
    The Ultimate Historical Timeline of Mechanical Lubrication
    Feb 4, 2019 · Between 1883 and 1905, new theories on hydrodynamic lubrication came to light. This was because extensive research showed that adding a ...
  8. [8]
    A Brief History of the Theory of Hydrodynamic Lubrication
    Sep 26, 2019 · Introduction. This paper offers a brief review of the history of the theory of hydrodynamic lubrication written on the occasion of what.<|control11|><|separator|>
  9. [9]
    Fluid Mechanics in Sommerfeld's School - Annual Reviews
    Apr 14, 2014 · In 1904, he published a theory of hydrodynamic lubrication. Four years later, he conceived an ap- proach for the analysis of flow instability ( ...
  10. [10]
    The history of the Stribeck curve and ball bearing steels
    May 12, 2010 · The historical development on the discoveries of lubrication regimes for lubricated sliding contacts has been reviewed from the very beginning.
  11. [11]
    Elastohydrodynamic lubrication | Friction and Wear in Engineering ...
    Concept introduced in the 1940s by Ertel and Grubin to explain unexpected film thickness in rolling contacts · Dowson and Higginson developed the first ...
  12. [12]
    Bearings (Machinery) -- Standards - The Online Books Page
    [X-Info] Analysis and lubrication of bearings (McGraw-Hill Book Co., 1949), by Milton C. Shaw and E. Fred Macks (page images at HathiTrust); [X-Info] ...
  13. [13]
    [PDF] Lubrication Background - NASA Technical Reports Server (NTRS)
    The present section attempts to trace the development of understanding of elastohydrodynamic lubrication by pointing out landmarks in the short history of this ...
  14. [14]
    Elastohydrodynamic Lubrication: A Gateway to Interfacial Mechanics ...
    Aug 9, 2025 · Elastohydrodynamic Lubrication (EHL) is commonly known as a mode of fluid-film lubrication in which the mechanism of hydrodynamic film ...
  15. [15]
    [PDF] Basic Lubrication Equations - NASA Technical Reports Server (NTRS)
    ... Osborne. Reynolds. Viscosity is the most important property of the luoricants employed in hydrQdynamic and elastohydrodynamic lubrication. In general ...
  16. [16]
    [PDF] HYDRODYNAMIC LUBRICATION THEORY
    A theoretical analysis of hydrodynamic lubrication was carried out by Osborne Reynolds. ... The following assumptions were made by Reynolds in the analysis:.Missing: scaling | Show results with:scaling
  17. [17]
    Long-scale evolution of thin liquid films | Rev. Mod. Phys.
    Jul 1, 1997 · In this review a unified mathematical theory is presented that takes advantage of the disparity of the length scales and is based on the asymptotic procedure.
  18. [18]
    [PDF] The propagation of two-dimensional and axisymmetric viscous ...
    viscous gravity currents over a rigid horizontal surface. By HERBERT E. HUPPERT. Department of Applied Mathematics and Theoretical Physics, Silver Street ...
  19. [19]
    [PDF] Wetting: statics and dynamics - PG de Gennes
    The wetting of solids by liquids is connected to physical chemistry (wettability), to statistical physics (pin- ning of the contact line, wetting transitions, ...
  20. [20]
    [PDF] The General Form of Reynolds Equation - MIT OpenCourseWare
    We combine with this with an integrated analysis of the flow in the vertical direction, and the use of what is commonly called the “kinematic boundary condition ...
  21. [21]
    [PDF] One-Dimensional Fluid Film Bearings A. Plane Slider Bearing B ...
    Figure 1 shows the geometry and coordinate system of a one-dimensional slider (thrust) bearing. The film thickness (h) has a linear taper along the direction of ...Missing: UL^ | Show results with:UL^
  22. [22]
    Performance enhancement of hydrodynamic thrust bearings
    Hydrodynamic bearings have gained prominence due to their high load-carrying capacity under extreme conditions. These bearings operate by generating a fluid ...
  23. [23]
    [PDF] FUNDAMENTALS OF FLUID SEALING John Lewis
    The fundamentals of fluid sealing, including seal operating regimes, are discussed. The general fluid-flow equations for fluid sealing are developed.Missing: gears | Show results with:gears
  24. [24]
    A Review on Gearbox Lubrication and Oil Leakages - AIP Publishing
    Mar 25, 2008 · This review covers gearbox lubrication, including hydrostatic, hydrodynamic, and elastohydrodynamic theories, and various techniques and ...
  25. [25]
    Lubrication Regime - an overview | ScienceDirect Topics
    According to the famous Stribeck curve [67,101] there are three lubrication regimes, namely: boundary; mixed and elastohydrodynamic (EHL); and hydrodynamic ...
  26. [26]
    [PDF] Engine Rotor Dynamics, Synchronous and Nonsynchronous Whirl ...
    In 1886 Reynolds derived his famous lubrication equation based on the theory that the resuliint pressure profile was due to hydrodynamic action in the fluid and ...
  27. [27]
    Journal Bearings and Their Lubrication
    High-speed journal bearings are always lubricated with oil rather than a grease. Oil is supplied to the bearing by either a pressurized oil pump system, an oil ...
  28. [28]
    [PDF] Hydrodynamic Journal Bearings - Dyrobes
    The relief track removes some of the bearing's load carrying capacity thereby forcing the bearing to operate at a higher eccentricity ratio. While adding a ...
  29. [29]
    I. On the theory of lubrication and its application to Mr. Beauchamp ...
    On the theory of lubrication and its application to Mr. Beauchamp tower's experiments, including an experimental determination of the viscosity of olive oil.
  30. [30]
    Research Report 3: Are Synovial Joints Squeeze-Film Lubricated?
    A simple analysis of squeeze films between compliant surfaces is presented and supported experimentally. Application of this analysis to synovial joints ...
  31. [31]
    The role of synovial fluid filtration by cartilage in lubrication of ...
    A mathematical model of lubrication of human synovial joints under squeeze-film conditions is presented in this several-part paper.Missing: permeable | Show results with:permeable
  32. [32]
    EXPERIMENTAL VERIFICATION OF THE ROLE OF INTERSTITIAL ...
    The study found that interstitial fluid pressurization is a primary mechanism in cartilage lubrication, reducing friction by supporting load transfer.
  33. [33]
    THE VISCOSITY OF THE BLOOD IN NARROW CAPILLARY TUBES
    Influence of Fahraeus-Lindqvist Effect on Blood Flow Resistance: An Analytical Approach. 1 Jan 2024. The flow of anisotropic nanoparticles in solution and in ...
  34. [34]
    The Fåhræus-Lindqvist effect in small blood vessels - PubMed Central
    According to it, in specific conditions a RBC-free layer develops near the vessel wall and this leads to a reduction of marginal fluid viscosity. Starting from ...
  35. [35]
    LUNG SURFACTANT: A Personal Perspective - Annual Reviews
    Mar 1, 1997 · Abstract This perspective tells the story of the discovery, characterization, and understanding of the surfactant system of the lung; ...
  36. [36]
    The Biophysical Function of Pulmonary Surfactant - PMC - NIH
    Pulmonary surfactant lowers surface tension in the lungs. Physiological studies indicate two key aspects of this function: that the surfactant film forms ...
  37. [37]
    Dip coating of cylinders with Newtonian fluids - ScienceDirect.com
    The Landau-Levich-Derjaguin (LLD) theory is widely applied to predict the film thickness in the dip-coating process.
  38. [38]
    The Thickness and Structure of Dip-Coated Polymer Films in ... - NIH
    The coating thickness was studied for a broad range of capillary numbers, 10−6 < Ca = U η / σ < 10−2, specifying the effect of the speed U of withdrawal of the ...Missing: blade lubrication
  39. [39]
    Thin-Film Flow at Moderate Reynolds Number | J. Fluids Eng.
    The main objective of this paper is to demonstrate the effectiveness of the Nusselt film equation for gravity-driven flow. Flow over a round-crested weir, ...
  40. [40]
    The effect of wall slip on the dewetting of ultrathin films on solid ...
    Nov 4, 2020 · The effect of wall slip on the dewetting of ultrathin films on solid substrates: Linear instability and second-order lubrication theory ...
  41. [41]
    Experimental Investigations of Biological Lubrication at the Nanoscale
    The aim of this review is to perform a critical analysis of the current stage of this research, with a main focus on studies on synovial joints and the oral ...
  42. [42]
    A Review of Thermal Effects in Hydrodynamic Bearings Part I
    Abstract. An extensive survey of the papers pertaining to thermal effects in hydrodynamic bearings, particularly the slider/thrust bearings, is presented ...Missing: seminal | Show results with:seminal
  43. [43]
    A Generalized Reynolds Equation for Non-Newtonian Thermal ...
    A new generalized Reynolds equation, which can incorporate most of the rheological laws found in the literature, is derived in this paper.
  44. [44]
    Power law fluid model incorporated into elastohydrodynamic ...
    We have considered a general solution including elastohydrodynamic lubrication (EHL) for different values of power law exponent. Deviation of values of central ...
  45. [45]
    Thermoelastohydrodynamic Analysis of Journal Bearings With Non ...
    Feb 16, 2010 · The analysis showed that the fluid characteristics as defined by the Carreau model, led to large differences in minimum film thickness and ...
  46. [46]
    [PDF] An Introduction To ElastoHydrodynamic Lubrication - Moodle
    As can be found from the equations in the section on the Hertzian contact, ph/a = E′/(πRx), thus: H(X, Y ) = H0 +. X2. 2. +. Y 2. 2. +. 2 π2 Z +∞. −∞ Z + ...
  47. [47]
    Effects of the Lubricant Piezo-Viscous Properties on EHL Line and ...
    Dec 13, 2012 · The EHL has two primary aspects: the increase of viscosity with pressure, and the elastic deformation, caused by high pressure, which is ...
  48. [48]
    Elastohydrodynamic Lubrication: Theory, Types & Practical Guide
    Elastohydrodynamic lubrication (EHL) describes a lubrication regime where high pressure causes significant elastic deformation of the contacting surfaces,History of Elastohydrodynamic... · Theory & Equations of...
  49. [49]
    Multigrid Methods for EHL Problems
    Sep 1, 1996 · In many bearings and contacts, forces are transmitted through thin continuous fluid films which separate two contacting elements.
  50. [50]
    An overview of functional biolubricants - SciOpen
    May 21, 2022 · Natural biomacromolecular lubricants are essential for maintaining ultra-low coefficients of friction between sliding biological interfaces.