Fact-checked by Grok 2 weeks ago

Friction

Friction is a that opposes the relative motion or tendency toward motion between two surfaces or systems in physical contact, acting parallel to their interface and always in the direction opposite to the impending or actual motion. This force arises from microscopic interactions, such as interlocking asperities and between surface atoms, and is crucial for enabling —such as walking, where static friction between shoes and ground propels the body forward—while also dissipating and causing wear in mechanical systems. Friction is categorized into two primary types: static friction, which acts on stationary objects to prevent the initiation of motion and can vary in magnitude up to a maximum value, and kinetic friction, which acts on moving objects to oppose sliding and remains roughly constant during uniform motion. The magnitude of frictional forces follows empirical laws primarily attributed to Guillaume Amontons and , though refined through modern observations. For both static and kinetic friction, the force is directly proportional to the normal force pressing the surfaces together, expressed as f_s \leq \mu_s N for static friction and f_k = \mu_k N for kinetic friction, where \mu_s and \mu_k are dimensionless coefficients of friction specific to the material pair, and N is the normal force. Typically, \mu_s > \mu_k, meaning it is easier to maintain motion than to start it, and kinetic friction is largely independent of sliding speed over typical ranges, though it does depend weakly on factors like and . These coefficients vary widely; for example, rubber on dry has \mu_s \approx 1.0 and \mu_k \approx 0.7, while polished on yields \mu_s \approx 0.6 and \mu_k \approx 0.3. Beyond these basics, friction plays a pivotal role in diverse fields, from engineering—where it influences tire grip, brake performance, and conveyor belt efficiency—to geophysics, such as in earthquake dynamics via stick-slip motion modeled by rate-and-state friction laws that account for velocity dependence and contact history. In biological systems, friction enables tactile sensation and muscle control, while in nanotechnology, it affects the operation of microelectromechanical systems (MEMS) at scales where atomic forces dominate. Reducing unwanted friction through lubricants or surface treatments is a key focus in tribology, the science of interacting surfaces in relative motion, which aims to minimize energy loss—estimated to contribute to approximately 23% of the world's total energy consumption.

Fundamentals

Definition and Scope

Friction is the tangential that acts to oppose the relative motion or tendency of relative motion between two surfaces in or between a surface and a surrounding medium. This arises at the where the bodies interact, acting parallel to that interface and in a direction that resists sliding or separation. In physics, friction—particularly kinetic friction—is a dissipative that converts into , whereas static friction often enables motion without dissipation; it plays a central role in both natural phenomena and human-engineered devices. At a high level, friction can be distinguished by the nature of interaction: contact friction, often termed or solid friction, occurs between solid surfaces through direct and molecular interactions, while non-contact friction encompasses types, involving viscous in liquids or gases, and electromagnetic types, which arise from magnetic or without physical touching. friction, for instance, manifests as drag on an object moving through air or , whereas electromagnetic friction appears in applications like magnetic braking systems where currents induce opposing forces. These distinctions highlight friction's broad scope beyond mere solid-solid contacts. Friction is crucial for enabling motion control in everyday activities and technologies, such as providing the grip necessary for walking on the ground or holding tools without slippage, which relies on static friction to prevent unintended movement. In engineered systems, it allows vehicles to accelerate, brake, and turn by transferring forces between tires and roads. Conversely, friction causes energy dissipation through heat generation and material wear, accounting for significant energy losses in machinery—estimated at about one-third of global energy resources—and necessitating designs that balance utility with efficiency. The occurrence of friction presupposes a perpendicular to the contact surface, which presses the bodies together, and either actual relative motion or an impending motion between them. As a vector quantity, the friction force has a magnitude measured in newtons () and a direction opposite to the relative velocity or attempted motion, ensuring it always acts to impede rather than assist the interaction.

Historical Overview

Observations of friction date back to ancient civilizations, where practical experiences with sliding and rolling objects informed feats like the construction of pyramids and the use of sledges lubricated with . However, systematic scientific investigation began in the with (1452–1519), who conducted pioneering experiments on friction in mechanical systems such as axles and screw threads. Through detailed sketches in his notebooks, da Vinci established that frictional resistance is proportional to the applied load and independent of the apparent contact area, insights derived from observing blocks sliding on inclined planes and rotating components. In the late 17th century, (1663–1705) independently rediscovered and experimentally validated these principles in his 1699 memoir presented to the Académie Royale des Sciences. Amontons' experiments with wooden blocks on inclined planes and pulleys demonstrated that the force required to initiate and maintain motion is roughly one-third of the normal load, confirming proportionality to load and independence from contact area or speed for dry surfaces. Building on this, (1736–1806) in his 1785 treatise "Mémoire sur un nouveau genre d'expérience pour déterminer les lois du frottement" refined the laws through precise measurements using torsion balances and inclined planes, notably introducing the critical distinction between static friction—which must be overcome to start motion—and kinetic friction, which opposes ongoing sliding. Coulomb's work also quantified that static friction exceeds kinetic friction, providing a foundational framework for applications. The 19th and early 20th centuries saw incremental refinements, including more precise distinctions between static and kinetic regimes and explorations of influencing factors like and material pairs. Advances in , a related distinct from sliding friction, gained prominence in the 1920s amid growing automotive and industries, with experimental methods like coastdown tests emerging to quantify losses in wheels. These studies emphasized in tire deformation as a primary cause, informing designs for reduced . The modern era of friction research accelerated in the 1960s with the Jost Report of 1966, commissioned by the British government, which coined the term "" to encompass , , and , estimating annual economic losses from poor tribological practices at hundreds of millions of pounds and advocating for interdisciplinary study. In the 1980s and 1990s, the advent of , including the atomic force microscope (AFM) invented in 1986, birthed nanotribology by enabling atomic-scale measurements of friction forces and revealing mechanisms like atomic stick-slip behavior. These tools uncovered deviations from classical laws at the nanoscale, such as velocity dependence and single-asperity contacts, profoundly influencing fields from to biomaterials.

Dry Friction

Basic Laws

The basic laws of dry friction, known as the Amontons-Coulomb laws, describe the frictional force between two dry surfaces in macroscopic contact as being proportional to the normal force pressing the surfaces together and independent of the apparent contact area. These empirical relations form the foundation of classical friction theory and were established through systematic experiments in the late 17th and 18th centuries. Guillaume Amontons derived the core principles in 1699 through pressure-based experiments involving sliding blocks of materials such as iron, , lead, and , often under light grease to simulate dry conditions. By varying the applied while keeping the contact area constant, Amontons observed that the frictional increased linearly with the normal , leading to the first law: the magnitude of the frictional force f is directly proportional to the normal force N, expressed as f = \mu N, where \mu is a constant coefficient of friction specific to the material pair. In further tests, he altered the apparent area—such as by changing block dimensions or using inclined surfaces—without observing a corresponding change in friction for a given normal force, establishing the second law: frictional force is independent of the apparent area. These findings built on earlier qualitative observations but provided the first quantitative empirical basis, assuming rough surfaces where was negligible. Charles-Augustin de Coulomb refined and validated these laws in the 1770s and 1780s using a torsion apparatus to measure sliding and static friction under controlled conditions. His experiments systematically varied parameters including , contact area, surface materials (e.g., wood, metal, brick), , and , confirming Amontons' proportionality and area independence across diverse setups. Coulomb's torsion involved suspending a weight via a twisted wire to quantify the required to initiate or maintain motion, revealing that friction arises primarily from the of surface asperities rather than cohesive forces, with the frictional force modeled as f = \mu N for sliding contacts. He also noted a small component in some cases but emphasized its minor role (<5% of total force) under typical dry conditions, solidifying the two-law framework. These laws apply specifically to dry, macroscopic contacts at low sliding speeds, where velocity dependence is negligible and surfaces are non-adhering with multi-asperity interactions; they neglect microscopic or high-speed effects. The frictional force always acts tangential to the surfaces and opposes the direction of impending relative motion, ensuring it either prevents sliding (static case) or resists ongoing motion (kinetic case). Experimental validation commonly employs setups, where the angle at which an object begins to slide yields \mu = \tan \theta, directly testing proportionality to (via weight component N = mg \cos \theta) and confirming area independence by comparing blocks of varying sizes. While the laws distinguish between maximum static friction and kinetic friction during sliding, their core formulation remains unified under f = \mu N.

Static and Kinetic Components

In dry friction, the static component refers to the frictional force that acts between two surfaces in contact when there is no relative motion between them. This force opposes the initiation of sliding and can vary in magnitude up to a maximum value given by f_s \leq \mu_s N, where \mu_s is the of static friction and N is the normal force pressing the surfaces together. The static friction adjusts dynamically to match the applied tangential force, preventing motion as long as this force remains below the maximum threshold, thereby maintaining in systems like a parked on an incline. Once relative motion begins, the frictional resistance transitions to the kinetic component, which provides a nearly constant opposing force during sliding, expressed as f_k = \mu_k N, where \mu_k is the of kinetic friction. Typically, \mu_k < \mu_s, meaning the force required to sustain motion is less than that needed to start it, a relationship observed across various material pairs and attributed to differences in at the interface. This distinction is evident in practical scenarios, such as brake pads that rely on static friction to hold a vehicle stationary but experience kinetic friction during sliding to decelerate it. The transition from static to kinetic friction occurs when the applied force exceeds the breakaway value of \mu_s N, initiating sliding. Factors like can influence this onset by altering the real area of contact and the interlocking of asperities, often increasing the static friction on rougher surfaces. In some systems, this transition leads to the stick-slip phenomenon, where motion alternates between static "stick" phases (building up ) and rapid kinetic "slip" phases (releasing that stress), resulting in or oscillations. For instance, a wooden block sliding on a rough table may exhibit stick-slip if pulled slowly, causing jerky motion, while smoother surfaces like lubricated metal sliders minimize this effect.

Coefficient and Normal Force

The normal force, denoted as N, is the component of the contact force between two objects that acts perpendicular to their interface at the point of contact. In vector terms, it is the force exerted by a surface to prevent interpenetration, directed orthogonally outward from the surface and balancing any perpendicular components of other forces, such as gravity. For an object on an inclined plane, the normal force often equals mg \cos \theta, where m is the mass, g is gravitational acceleration, and \theta is the incline angle, assuming no acceleration perpendicular to the surface. The coefficient of friction, \mu, is a dimensionless empirical defined as the of the frictional f to the normal N, such that f = \mu N. It quantifies the frictional between two surfaces and is inherently material-dependent, varying based on the specific pair of contacting materials. Separate coefficients exist for static friction (\mu_s), which opposes impending motion, and kinetic friction (\mu_k), which acts during sliding, with \mu_s typically greater than \mu_k. According to Amontons' laws, the frictional (and thus \mu) is independent of the apparent contact area, depending only on N and the materials involved. Several factors influence the value of \mu, including , which can alter surface properties and effects, and , such as roughness or cleanliness, which affects asperity interactions. Despite these variations, the basic model assumes \mu remains constant for given conditions, as an empirical approximation. Typical coefficients for common material pairs are summarized in the following table, illustrating approximate ranges under dry conditions:
Material Pair\mu_s (Static)\mu_k (Kinetic)
Rubber on concrete0.6–0.850.45–0.75
Steel on steel (dry)0.5–0.80.42
Ice on ice0.10.02
These values are representative and can vary with specific conditions like moisture or contaminants. Coefficients of friction are measured using techniques such as direct force sensors, where a tangential force is applied via a probe until sliding occurs (for \mu_k) or impending motion (for \mu_s), with N controlled separately. Alternatively, the tilt method involves inclining a surface until sliding begins, where \mu_s = \tan \theta at the critical angle, and N is derived from the setup geometry.

Angle of Friction

The angle of friction, denoted as φ, is defined as the angle between the resultant reaction force (comprising the normal force and the frictional force) and the normal force acting on a surface at the onset of sliding in dry friction scenarios. This angle provides a geometric interpretation of the coefficient of friction μ, where the relationship is given by: \tan \phi = \mu Here, φ represents the critical angle at which the frictional force exactly balances the component of gravity tending to cause motion down an incline, assuming static equilibrium just before sliding occurs. For instance, on an inclined plane, if the incline angle θ equals φ, the object is on the verge of slipping, illustrating how friction prevents motion under gravitational pull. The angle of friction can be experimentally determined using a tilting plane apparatus, where a is placed on a surface that is gradually inclined until the block begins to slide. The angle θ at which sliding initiates is measured, and μ is calculated as tan θ, with φ thus equaling θ for the static case. This method is straightforward and widely used in laboratory settings to quantify frictional properties between solid surfaces. In the context of granular materials, the angle of repose θ is the maximum inclination angle at which a pile of loose particles remains stable without collapsing, and it corresponds directly to the angle of friction φ under ideal dry conditions. For static equilibrium, θ = φ, as the interparticle friction balances the gravitational component along the slope. This concept is applied in for designing storage piles, hoppers, and prediction, where the repose angle influences material flow and stability. The internal friction angle, also denoted φ, extends this concept to and , representing the material's resistance to due to particles in cohesionless soils like sands. In the simplified Mohr-Coulomb failure criterion for such materials (where cohesion c is negligible), the τ relates to the normal σ by: \tau = \sigma \tan \phi Typical values for loose sands range from approximately 30° to 35°, highlighting how particle angularity and density affect resistance. This angle is crucial for analyzing and foundation design in granular deposits. Visually, the angle of friction is represented in the cone of friction diagram (sometimes called the pyramid of forces in two dimensions), where the possible directions of the resultant force lie within a (or ) whose semi-vertical is φ, centered on the normal force vector. This geometric construct illustrates that any applied force within the cone does not overcome static friction, while forces outside it cause impending motion.

Atomic-Scale Mechanisms

At the atomic scale, friction originates from interactions at asperity contacts between surfaces, where the real area of contact is much smaller than the apparent area due to . between these asperities, primarily driven by van der Waals forces, leads to the formation of junctions that must be sheared during sliding, contributing significantly to the frictional force. Additionally, plowing occurs when harder asperities deform or displace softer , generating a deformation component to friction that depends on the and difference of the contacting surfaces. In single-asperity models, which idealize contact at isolated nanoscale protrusions, the frictional force f is proportional to the real contact area A_r and the G of the material, expressed as f \propto G A_r. This relationship arises because the interfacial is typically a fraction of the material's shear modulus, reflecting the required to break bonds or slide planes within the . Such models bridge atomic interactions to measurable forces, emphasizing that friction scales with the actual interacting area rather than nominal . The Prandtl-Tomlinson model provides a foundational description of atomic-scale friction, portraying the sliding tip as a particle in a combined potential: a periodic lattice potential from the substrate and a parabolic confining potential from the support. As the support moves, the tip experiences stick-slip behavior, remaining pinned in potential minima (stick) until the force exceeds a threshold, causing abrupt slips to the next minimum. This model captures the discrete nature of atomic motion and explains the sawtooth-like friction traces observed experimentally. At the nanoscale, friction exhibits velocity dependence, with higher sliding speeds reducing stick-slip amplitude due to inertial effects or thermal activation overcoming barriers more readily. Experimental investigations using (AFM) have been instrumental in probing these mechanisms, enabling precise measurements of friction on well-defined surfaces like or . AFM studies reveal that the friction coefficient \mu often varies nonlinearly with applied load at the atomic scale, frequently decreasing as load increases due to changes in contact area and dominance at low loads. For instance, on clean surfaces, \mu can drop from values exceeding 0.1 at ultralow loads to near macroscopic levels at higher loads, highlighting the transition from adhesion-controlled to deformation-controlled regimes. In modern contexts, quantum effects introduce additional nuances to atomic-scale friction, particularly in superconductors where electron-phonon interactions modulate energy dissipation. Below the superconducting transition temperature, friction can decrease markedly as pairs suppress electron-phonon , reducing drag on sliding asperities; AFM experiments on high-Tc superconductors such as Bi-2212 show approximately 30% lower friction forces in the superconducting state compared to the normal state, while for NbSe₂, reductions by a of up to 3 have been observed. These observations underscore the role of electronic in nanoscale .

Model Limitations and Instabilities

The classical model of dry friction, which assumes a constant of friction independent of , area, and , exhibits significant limitations under certain conditions. At high sliding speeds, friction often shows dependence, where the decreases due to effects or viscoelastic responses in the interface, deviating from the model's prediction of constancy. In adhesive s, such as those involving polymers or soft materials, friction force can depend on the real area rather than solely on the normal load, leading to non-Amontons behavior observed in experiments with varying and strength. variations further challenge the model; elevated temperatures can cause lubricants to melt or degrade, reducing and altering the friction from to hydrodynamic, as seen in solid lubricants like metals where friction increases or decreases at melting points. A notable exception is the phenomenon of superlubricity, where the effective friction coefficient can become negative, meaning friction decreases with increasing normal load due to structural incommensurability at the interface. This occurs in layered materials like sliding against hexagonal , where interlayer collapses under load, achieving coefficients as low as 10^{-4} without . Such "negative" coefficients highlight the model's failure to capture load-dependent weakening in atomically smooth contacts. Dynamic instabilities in dry friction systems further reveal the Coulomb model's inadequacies, as it does not account for nonlinear behaviors like chaos and bifurcations. Chaotic motion arises in forced oscillators with dry friction, where stick-slip transitions lead to unpredictable trajectories, as demonstrated in experimental spring-mass setups with harmonic forcing. Bifurcations can trigger self-excited vibrations, such as squealing in brakes from mode coupling or earthquake ruptures from velocity-weakening on faults, where steady sliding becomes unstable. To address these limitations, numerical simulations using finite element methods incorporate the model with dynamic equations, such as the for a sliding block m \frac{da}{dt} = -\mu m g \operatorname{sign}(v), solved alongside constraints to predict instabilities in complex geometries. More advanced rate-and-state friction laws extend the model for seismic applications by including and evolution, given by \mu = \mu_0 + a \ln\left(\frac{V}{V_0}\right) + b \ln\left(\frac{\theta V_0}{D_c}\right), where \mu_0 is the reference friction, a and b are material parameters, V is slip , \theta is a representing contact age, V_0 is reference , and D_c is critical slip distance; this formulation captures velocity-weakening leading to dynamic instabilities like aftershocks.

Fluid and Boundary Friction

Fluid Friction Principles

Fluid friction, also known as viscous , arises from the internal resistance within a medium when layers of the fluid move relative to one another, opposing the motion of an object through the fluid. Unlike dry friction, which is independent of velocity and depends on , viscous is directly proportional to the relative between the object and the fluid. The force f_d can be expressed as f_d = -b v, where b is the damping coefficient and v is the ; this linear relationship holds for low-speed flows where inertial effects are negligible. The foundational principle governing fluid friction is Newton's law of viscosity, which states that the \tau in a is proportional to the du/dy between adjacent layers: \tau = \eta \frac{du}{dy}, where \eta is the dynamic of the , a measure of its resistance to shear. This law applies to Newtonian fluids, where remains constant regardless of the . Dynamic \eta has units of ·s and depends on the 's molecular structure and , with denser fluids like oils exhibiting higher values than gases. In practical terms, this generates the frictional force that slows objects moving through the , such as a sedimenting in a column. Fluid friction manifests differently based on flow regimes: , characterized by smooth, parallel layers with minimal mixing, and turbulent flow, marked by chaotic eddies and enhanced momentum transfer to the flow direction. The transition between these regimes is determined by the Re = \frac{\rho v L}{\eta}, where \rho is and L is a ; low Re (typically below 2000) favors with lower friction, while high Re leads to and increased . For small spheres in at low s, provides the viscous force as f = 6 \pi \eta r v, where r is the sphere's ; this , derived from solving the Navier-Stokes equations under creeping flow assumptions, accurately predicts velocities in applications like particle settling in air or . In more complex scenarios, such as objects moving at higher speeds, effects dominate friction through skin friction, the tangential force exerted by the on the object's surface. The is a thin region near the surface where gradients are steepest, leading to significant . The total drag force f relates to the skin friction C_d via C_d = \frac{f}{0.5 \rho v^2 A}, where A is the reference area; for flat plates in , C_d scales inversely with \sqrt{Re}, while turbulent boundary layers yield higher but more gradually decreasing coefficients. Unlike solid-solid friction, friction does not involve a component but is governed by properties like \rho and \eta, influencing applications in and hydrodynamics where minimizing drag is critical.

Lubricated and Boundary Regimes

In lubricated contacts, friction behavior varies across distinct regimes depending on the thickness of the film relative to . Hydrodynamic lubrication occurs when a full fluid film completely separates the contacting surfaces, preventing direct solid-solid interaction. The frictional force in this regime arises primarily from viscous shear in the and can be approximated as f \propto \eta U L^2 / h, where \eta is the viscosity, U is the sliding speed, L is the length of the contact, and h is the film thickness. This regime is governed by the , which describes pressure generation in thin films due to surface motion. Under high loads, such as in or rolling-element bearings, elastic deformation of the surfaces becomes significant, leading to elastohydrodynamic lubrication (EHD). In EHD, the lubricant film thickness remains thin (often on the order of nanometers) but sufficient to support the load through combined hydrodynamic and elastic compliance, differing from pure hydrodynamic lubrication by accounting for surface deflection and pressure-induced viscosity increases. This regime is essential for high-pressure applications where rigid-body assumptions fail. Boundary lubrication emerges when the lubricant film is thinner than the composite , resulting in direct asperity-asperity contact that bears most of the load. Friction here is dominated by plowing of asperities or entrapped debris, with coefficients typically ranging from 0.01 to 0.1, reduced by additives that form protective layers on surfaces. The transitions between these regimes are captured by the Stribeck curve, which plots the friction coefficient \mu against the Hersey number \eta U / P (where P is the mean ). At low Hersey numbers, boundary lubrication prevails with high \mu; as the number increases, \mu decreases through the mixed regime before rising in the hydrodynamic regime due to viscous effects. The mixed regime combines elements of boundary and hydrodynamic lubrication, where the load is shared between fluid film and asperity contacts. Friction in this transitional state is modeled by partitioning contributions from solid interactions and viscous shear, often using or deterministic approaches that incorporate statistics. Lubrication films can fail through mechanisms like , where insufficient lubricant supply thins the film and promotes asperity contact, or , where vapor bubbles form and collapse in low-pressure regions, disrupting film continuity and leading to localized . These modes are particularly relevant in mixed or boundary conditions under varying operating speeds or loads.

Specialized Friction Phenomena

Skin and Internal Friction

Skin friction refers to the resistive force experienced by human skin or similar surfaces when moving through fluids, such as water during swimming, where it contributes to overall drag alongside pressure and wave components. In underwater gliding, the total drag coefficient for a human swimmer in a prone position can reach approximately 1.0 at typical speeds like 1.6 m/s, with skin friction accounting for about 7-16% of the total drag depending on body position and velocity. This fluid-induced skin friction arises from shear stresses at the skin-water interface, influenced by surface roughness and boundary layer effects, and can be briefly contextualized within broader fluid friction principles where viscous forces dominate near the surface. Beyond fluid interactions, skin friction plays a key role in scenarios, such as gripping objects, where the of friction for dry on surfaces averages around 0.6 for the , enabling secure handling without slippage. This value reflects the interplay of skin's viscoelastic and surface , which enhance tangential under loads. Internal friction, also known as material , describes the within or tissues during deformation, quantified by the loss tangent tan δ, defined as the ratio of energy dissipated per to the maximum stored. High tan δ values indicate greater , where frictional losses convert into through internal mechanisms like molecular sliding. Viscoelastic materials exhibit this behavior, modeled by elements such as the model, which combines elastic and viscous responses via the : \sigma + \tau \frac{d\sigma}{dt} = \eta \frac{d\varepsilon}{dt} Here, σ is stress, ε is strain, τ is the relaxation time (η/G, with G as the elastic modulus), and η is viscosity; this equation captures time-dependent relaxation where stress decays under constant strain due to internal frictional flows. In polymers, internal friction often manifests as hysteresis, arising from energy losses during cyclic loading as molecular chains rearrange and slide past one another, leading to a looped stress-strain curve where the area represents dissipated work. This molecular-level friction contributes to damping in applications like vibration absorption, with hysteresis proportional to the extent of chain entanglement and relaxation dynamics. Biological tissues exemplify internal friction reduction through specialized ; in synovial joints, , enriched with lubricin (proteoglycan 4), lowers the of friction between surfaces to below 0.01, minimizing energy loss and wear by forming protective boundary layers that counteract viscous and adhesive dissipation. This mechanism ensures efficient motion while preserving joint integrity against internal frictional heating. Dynamic mechanical analysis (DMA) measures internal friction by applying oscillatory loads and tracking tan δ as a function of temperature or frequency, revealing damping peaks at transitions like the glass-rubber state in polymers or tissues.

Rolling Resistance

Rolling resistance is the force that opposes the motion of a rolling body on a surface, arising primarily from energy dissipation in the deformation of the contacting materials rather than from sliding at the interface. This force, denoted as f_r, is commonly expressed as f_r = C_r N, where N is the normal force and C_r is the dimensionless coefficient of rolling resistance, which typically ranges from approximately 0.001 to 0.03 depending on the materials involved. For instance, rubber tires exhibit higher C_r values (around 0.01 to 0.03) due to their viscoelastic properties, while steel wheels on steel rails show lower values (about 0.001 to 0.002) owing to minimal deformation. The primary mechanism of rolling resistance is elastic hysteresis, where energy is lost as heat during the cyclic deformation and recovery of the materials in the contact zone. In deformable materials like rubber tires, the hysteresis loop in the stress-strain curve leads to incomplete energy recovery, with the forward creep of the contact patch causing compression ahead of the roller and expansion behind it, resulting in net energy dissipation. For harder materials such as steel wheels on rails, contributions include both hysteresis from elastic deformation of the wheel and rail under load, as well as adhesion effects where molecular bonds at the interface require work to break and reform during rolling. A simplified derivation of the rolling resistance relates it to geometric deformation parameters: f_r = \frac{b}{r} N, where b is the characteristic deformation distance (such as the horizontal offset due to sinking or in the ) and r is the of the rolling body. This arises from the balance, where the frictional acts at a arm approximately equal to b, leading to C_r = \frac{b}{r}, thus linking the directly to the ratio of deformation to . Influencing factors include material properties, with softer, more hysteretic materials like rubber yielding higher C_r compared to rigid ones like ; speed, which causes a slight increase in resistance due to viscoelastic rate dependence; and , which modulates losses—typically reducing C_r for rubber as rises and tan δ (loss tangent) decreases. Unlike sliding friction, occurs in pure rolling without slip, where there is no relative tangential velocity at the instantaneous between the rolling body and surface, distinguishing it from friction regimes that involve direct shearing.

and Triboelectric Effects

friction, also known as , arises when a accelerates and emits , resulting in a self-force that opposes the motion. This phenomenon is described by the Abraham-Lorentz formula, which in non-relativistic cgs units gives the force as \mathbf{F}_\text{rad} = \frac{2}{3} \frac{q^2 \dot{\mathbf{a}}}{c^3}, where q is the charge, \dot{\mathbf{a}} is (time derivative of ), and c is the . In relativistic contexts, the Abraham-Lorentz-Dirac equation extends this to include velocity-dependent terms, accounting for the on the particle. This self-force plays a critical role in , where it represents the interaction of a with its own , influencing dynamics in high-energy environments like accelerators. The involves the transfer of between two contacting surfaces, typically insulators, during friction or contact-separation, leading to buildup. This charge transfer occurs primarily through exchange at the , where materials with differing electron affinities gain or lose s upon . The triboelectric series ranks materials by their tendency to become positively or negatively charged; for example, tends to charge positively while Teflon charges negatively when rubbed together. Mechanisms include electron tunneling or breakdown at asperities, where localized exceed the material's breakdown strength, facilitating charge flow. The resulting charge Q on a surface leads to a voltage buildup given by V = Q / C, where C is the , potentially reaching thousands of volts and causing electrostatic attraction or discharge. Applications of the include everyday static cling, where charged fabrics adhere to neutral surfaces due to electrostatic forces. In , however, tribocharging contributes to wear and failure by inducing that damages sensitive components or increases frictional . In modern , tribocharging significantly alters friction coefficients; for instance, electrostatic charges generated during sliding can increase the effective friction force by up to an order of magnitude in nanostructures, beyond contributions from . As of 2025, the has enabled triboelectric nanogenerators (TENGs) for harvesting from motion, powering wearable devices and sensors in sustainable applications.

Belt and Braking Friction

Belt friction arises in systems where a flexible or wraps around a cylindrical or , such as in conveyor systems or capstans, where the frictional interaction allows a small input to support a much larger load due to the amplification from the contact arc. The relationship is governed by the , also known as the Euler-Eytelwein formula: \frac{T_2}{T_1} = e^{\mu \theta}, where T_2 is the on the tight side, T_1 is the on the slack side, \mu is the coefficient of kinetic friction, and \theta is the wrap angle in radians. This equation originates from early 19th-century and has been widely validated in applications for predicting transmission in wrapped belts. The derivation of the capstan equation considers the belt as composed of infinitesimal elements over the wrap angle. For a small angular element d\theta, the normal force on the element is T d\theta, and the frictional force opposing slip is \mu T d\theta, leading to a differential change in tension dT = \mu T d\theta. Integrating this separable differential equation from \theta = 0 (slack side) to \theta (tight side) yields \ln(T_2 / T_1) = \mu \theta, or equivalently the exponential form. This model assumes constant \mu, no belt slip, and negligible belt mass or elasticity, making it applicable to steady-state operations in conveyor belts or mooring ropes. In braking systems, friction between pads and rotors or drums dissipates as , with the total energy E = \mu N d where N is force and d is the sliding distance, equivalent to heat generation Q = \mu N v t for sliding v and time t. Prolonged or high-energy braking causes temperatures to rise significantly, often exceeding 500°C in disk brakes, leading to where the effective \mu decreases due to material degradation, gas formation at the interface, or altering contact. Sintered metal brake pads, composed of compacted metal powders like and iron, provide stable performance with \mu \approx 0.4 under high loads and temperatures up to 600°C, outperforming organic materials in heavy-duty applications. Braking instabilities include , where rising temperatures accelerate wear and reduce \mu, increasing slip and further heat buildup in a loop, and vibrations such as squeal or judder in disk brakes arising from frictional self-excitation and modal coupling between pad and . These phenomena can limit braking efficiency, with vibrations often occurring above 100 Hz due to stick-slip dynamics at the . involves and cooling designs to prevent such escalations.

Friction Reduction

Mechanical Devices

Mechanical devices play a crucial role in reducing friction through physical engineering principles, such as minimizing contact areas, employing non-contact support, and optimizing load distribution, thereby enhancing efficiency in machinery without relying on chemical additives. Bearings represent a foundational class of mechanical devices designed to minimize sliding friction by converting it into rolling motion. Ball bearings achieve this by using spherical elements that establish point contacts between the rolling balls and raceways, significantly reducing the frictional area compared to plain sliding surfaces. Roller bearings, in contrast, employ cylindrical or tapered rollers that create line contacts, distributing loads more evenly along the length of the rollers while still limiting friction to rolling rather than sliding. These configurations yield coefficients of rolling friction typically below 0.0015 for ball bearings and around 0.001 for optimized roller types under normal operating conditions, enabling smoother rotation and lower energy losses in applications like motors and conveyor systems. Air and gas bearings further advance friction reduction by establishing a pressurized that completely separates moving surfaces, approaching near-zero mechanical contact. In these devices, or gas is supplied through porous media or nozzles to form a thin, stable —often on the order of micrometers thick—between the bearing surfaces, supporting loads via hydrodynamic pressure without physical touching. This results in friction coefficients effectively approaching zero, as the only resistance arises from minimal viscous in the gas , making air bearings ideal for high-precision machines such as coordinate measuring devices and manufacturing equipment where sub-micrometer accuracy is required. The inherent from the squeeze also provides , enhancing under dynamic loads. Magnetic levitation systems eliminate mechanical friction entirely by suspending components without physical contact, using electromagnetic forces for support and guidance. In these devices, superconducting or permanent magnets generate repulsive or attractive fields that levitate a or above a or , preventing any surface interaction and thus removing sliding or rolling friction. This no-contact approach drastically reduces dissipation, with applications like trains achieving speeds over 500 km/h while minimizing wear and maintenance needs due to the absence of traditional frictional losses. Geometry optimizations in mechanical components enhance friction reduction by promoting uniform load distribution, which prevents localized high- contacts that amplify and sliding. Tapered rollers in bearings, for instance, feature conical profiles that align axial and radial loads along converging lines, ensuring even across the length and minimizing edge overloading. Similarly, crowned incorporate a slight barrel-shaped on tooth flanks to compensate for misalignment and deflection, shifting load toward the tooth center and reducing peak stresses that could induce frictional heating or slip. These modifications improve load distribution and reduce effective friction in high-load scenarios by avoiding uneven patterns. Active control systems integrate sensors and actuators to dynamically mitigate friction-induced vibrations, adapting in real-time to varying conditions. These devices employ accelerometers or strain gauges to detect oscillatory motions caused by frictional instabilities, then use piezoelectric or electromagnetic actuators to apply counter-forces that resonances and stabilize contact interfaces. In friction damping applications, such as bolted joints or sliding mechanisms, active control adjusts clamping forces to optimize loads, reducing relative slip and associated energy losses while preventing . This approach achieves substantial reductions in amplitudes, indirectly lowering effective friction in precision assemblies like robotic arms.

Chemical Lubricants

Chemical lubricants play a crucial role in reducing friction by forming thin films that separate contacting surfaces, particularly in boundary lubrication where direct asperity contact is minimized through adsorbed layers. These lubricants consist primarily of base oils enhanced with additives to optimize performance under varying conditions such as temperature, pressure, and load. The primary types of chemical lubricants include oils, synthetic oils, and greases. oils, derived from refined base stocks, are widely used due to their cost-effectiveness and adequate in moderate conditions. Synthetic lubricants, such as polyalphaolefin (), offer superior performance in high-temperature environments by providing better thermal stability and oxidation resistance compared to oils. Greases are semi-solid formulations created by thickening base oils—often or synthetic—with soaps or non-soap agents like or , enabling them to stay in place on vertical surfaces and deliver sustained lubrication. To enhance friction reduction and , lubricants incorporate specialized additives. Anti- additives, such as dialkyldithiophosphate (ZDDP), react under load to form durable tribofilms on metal surfaces via , preventing severe in conditions. Friction modifiers, including (MoS₂) or organomolybdenum compounds like molybdenum dialkyldithiocarbamate (MoDTC), deposit low--strength layers that can achieve coefficients of friction below 0.05, significantly lowering energy losses. These films, formed through of additive molecules onto surface asperities, create sacrificial layers that reduce direct metal-to-metal contact and resistance. Lubricant selection emphasizes the (VI), a measure of how (η) remains stable across ranges; high-VI fluids, often synthetics, exhibit minimal changes, ensuring consistent thickness from low to high temperatures. Engineers select using standardized viscosity-temperature charts, such as those based on ASTM D341, to match operational requirements like ISO viscosity grades. Post-2000 environmental regulations, including the U.S. EPA's 2013 Vessel General Permit (VGP) and the 2024 Vessel Incidental Discharge Act (VIDA) standards, have driven the development of biodegradable lubricants—typically oil-based or synthetic esters that degrade over 60% within 28 days—to reduce ecological persistence in sensitive areas like waterways.

Practical Applications

Transportation Systems

In transportation systems, friction plays a pivotal role in vehicle , , and , influencing components from ground contact to internal mechanisms. Tires, for instance, rely on carefully engineered tread patterns to balance traction and . These designs channel water away during wet conditions to maintain , achieving friction coefficients of approximately 0.5 to 0.6 for effective wet-road in optimized treads. However, in tires—arising from in the rubber compound and deformation—contributes substantially to energy loss, accounting for 20-30% of a vehicle's total consumption depending on speed, load, and conditions. Braking systems harness controlled friction to decelerate vehicles safely. Disk brakes, which use caliper-mounted pads to clamp a rotating , provide superior dissipation and consistent performance under repeated or high-intensity use compared to drum brakes, where shoes expand against an enclosed interior, potentially leading to fade from trapped and gases. In electric vehicles, integrates motor-generated electromagnetic forces to recapture during slowing, reducing dependence on traditional friction brakes and cutting wear on pads and by 64-95% in typical driving scenarios. Within engines, friction at the ring-cylinder interface represents a major efficiency hurdle. Lubrication regimes, often or mixed, minimize direct metal-to-metal contact, yet these losses still consume 10-20% of the engine's indicated power, with rings alone responsible for up to 20% of total mechanical friction. Aerodynamic effects introduce , the viscous shear along vehicle surfaces, which becomes dominant at high speeds. For in cruise, skin friction accounts for roughly 50% of total drag, while in automobiles at highway velocities exceeding 100 km/h, overall aerodynamic drag—including skin friction components—can comprise up to 80% of resistance, underscoring the need for smooth contours and management. Advancements in electric vehicles incorporate low-friction (low-μ) coatings on elements like and bearings to curb parasitic losses, enhancing range and durability by reducing wear under high-torque conditions. Globally, friction-related inefficiencies in exact a heavy toll, with losses equivalent to about 33% of use in heavy-duty vehicles such as trucks and buses.

Measurement Methods

Direct methods for measuring friction include the technique, which determines the static of friction by gradually increasing the inclination angle until an object begins to slide, corresponding to the angle of repose where the frictional force balances the component of . This method is straightforward and widely used for bulk materials, providing a direct assessment of static friction under gravitational loading. Tribometers represent another primary direct approach, particularly the pin-on-disk configuration, where a stationary pin slides against a rotating disk to quantify the kinetic of friction (μ) through force sensors measuring tangential and normal loads. In this setup, the friction is calculated as the ratio of frictional force to normal load, enabling evaluation of and friction under controlled sliding conditions. Standardized protocols, such as ASTM G99, govern pin-on-disk testing to ensure reproducibility in assessments of and friction for materials like metals, coatings, and lubricants. This standard specifies parameters like load, speed, and duration, facilitating comparisons across studies while emphasizing the need to report all deviations from the procedure. Indirect methods involve dynamometers, which measure braking friction in automotive applications by simulating vehicle inertia and recording or during deceleration. These devices assess friction couple behavior under dynamic conditions, such as effects, providing insights into performance without direct contact isolation. At the nanoscale, (AFM) indirectly evaluates friction by raster-scanning a tip over a surface while monitoring lateral forces as a of load, revealing atomic-scale variations in frictional response. Advanced techniques like measure film thickness, which inversely relates to friction levels in elastohydrodynamic contacts by visualizing interference fringes from light reflected off the layer. This method achieves nanometer resolution for thin films, aiding in the study of boundary regimes where friction transitions from hydrodynamic to mixed. Key error sources in friction measurements include misalignment of load axes, which can introduce between normal and tangential forces, leading to inaccuracies up to several percent in low-friction scenarios. Surface , such as or lubricants, alters contact conditions and reduces measured friction by introducing hydrodynamic effects or altering . Typical for of friction determinations reaches 0.01 in well-calibrated setups, though this varies with and environmental controls.

Everyday and Industrial Uses

In everyday applications, friction plays a crucial role in writing instruments like pencils, where the core deposits material onto through controlled frictional contact. The low of kinetic friction between graphite and paper enables smooth writing by shearing off fine graphite particles without excessive resistance. Similarly, safety matches rely on high frictional interaction between the match head and the striking surface to generate localized , often exceeding 500°C, which ignites the chemical mixture of and red phosphorus. In industrial settings, friction is managed during drilling operations using twist bits, where lubricants such as nano-enhanced cutting oils reduce the coefficient of friction between the bit and workpiece, minimizing heat buildup and extending life by up to 50% in . Conveyor belts utilize the capstan effect, described by the Euler-Eytelwein , to amplify through frictional wrapping around , allowing efficient material transport with coefficients of friction between belt lagging and pulley typically around 0.3 to 0.5. Friction provides essential positive functions in fasteners and closures. Self-locking screws maintain position without additional components when the thread lead angle is less than the friction angle, determined by the coefficient of friction (often 0.1 to 0.2 for lubricated threads), preventing unintended rotation under load. In clothing zippers, controlled friction between , teeth, and ensures secure during , enabling smooth operation while resisting accidental separation. Despite these benefits, friction poses challenges through in tools and machinery, contributing to global annual economic losses estimated at approximately 2.5 trillion euros from material degradation and across industries as of 2018. accumulation exacerbates this by increasing the of friction on surfaces, with light dust layers causing up to a tenfold rise in some engineering applications, leading to higher and accelerated . Innovations in self-cleaning surfaces address these issues by mimicking lotus leaf microstructures to repel contaminants, thereby reducing and dirt accumulation on appliance exteriors and lowering maintenance needs in household devices like refrigerators and washing machines.

References

  1. [1]
    5.1 Friction – College Physics - University of Iowa Pressbooks
    Friction is a force opposing relative motion between systems in contact. It includes static friction when stationary and kinetic friction when moving.
  2. [2]
    Friction - Physics
    Sep 24, 1999 · Friction is a force parallel to the interface between objects, opposing relative motion. It can be static when no motion occurs, or kinetic ...
  3. [3]
    6.2 Friction – General Physics Using Calculus I - UCF Pressbooks
    Friction is a force opposing relative motion between systems in contact. Static friction occurs when stationary, and kinetic friction when moving.
  4. [4]
    Friction – Introductory Physics for the Health and Life Sciences I
    Friction is a force opposing relative motion between surfaces. It has two types: static, which prevents motion, and kinetic, which acts when objects slide.
  5. [5]
    [PDF] 6. Friction
    Friction is a force between surfaces, opposing relative motion. It has two types: kinetic (when surfaces move) and static (when surfaces don't move).
  6. [6]
    Content (Friction) | BoxSand – Flip the Classroom
    Friction is a force that arises between two objects that have surfaces in contact with one another. Friction opposes the relative motion between two objects ...
  7. [7]
    Friction - Summary - The Physics Hypertextbook
    "Friction" is often synonymous with "dry friction". Viscous friction. The resistive force between surfaces in relative motion through a fluid (liquids & gases).
  8. [8]
    Friction without contact | Nature Physics
    More surprising, friction can exist even between bodies that make no contact whatsoever. For example, measurements of the force required to drag the aluminium ...
  9. [9]
    Magnetic non-contact friction from domain wall dynamics ... - arXiv
    Sep 19, 2018 · Magnetic friction is a form of non-contact friction arising from the dissipation of energy in a magnet due to spin reorientation in a magnetic ...
  10. [10]
    The Physics Behind Friction - JHU Engineering Magazine
    Without the force called friction, cars would skid off the roadway and objects would tumble off tables and onto the floor. Even so, how friction works at a ...
  11. [11]
    New Law of Physics Helps Humans and Robots Grasp the Friction of ...
    Apr 29, 2021 · One reason friction is important is because it helps us hold things without dropping them. ... importance for engineering the mechanics of soft ...
  12. [12]
    [PDF] On the thermodynamics of friction and wear-A review
    Apr 27, 2010 · As the principal cause of wear and energy dissipation, friction is responsible for consuming about one-third of the world's energy resources as.
  13. [13]
    History of science friction - TRIBOLOGY-ABC
    Leonardo Da Vinci (1452-1519) was one of the first scholars to study friction systematically. He realized how important friction is for the workings of machines ...
  14. [14]
    A note on Guillaume Amontons and the laws of friction - Sage Journals
    Aug 25, 2021 · The widely cited paper of December 1699, in which Amontons restated his laws of friction more forcefully and explicitly,5 describes his further ...
  15. [15]
    Amontons Law - an overview | ScienceDirect Topics
    Leonardo da Vinci first carried out systematic investigations on the effects of friction about 500 years back in the 15th century (about 150 years before the ...
  16. [16]
    Kinetic Friction - an overview | ScienceDirect Topics
    Kinetic friction is defined as the force that opposes the motion of a body, acting in the direction opposite to its movement. The magnitude of kinetic ...Missing: 20th | Show results with:20th
  17. [17]
    [PDF] Historical Perspectives On Tire Rolling Resistance - Rubber News
    Sep 22, 2010 · This background provides historical context and perspective for current interest in Tire Rolling Resistance.
  18. [18]
    [PDF] Rolling Resistance – Basic Information and State-of - DiVA portal
    The rolling resistance coefficient (RRC) was first developed to present a characterization of ... Coastdown measurements were done as early as in the 1920's ...<|separator|>
  19. [19]
    The Invention of Tribology: Peter Jost's Contribution - MDPI
    Feb 21, 2024 · Peter Jost, known as the father of tribology, promoted the subject, authored the 'Jost Report', and established the International Tribology ...
  20. [20]
    Scratching the Surface: Fundamental Investigations of Tribology with ...
    A few years after the invention of the scanning tunneling microscope (STM), the atomic force microscope (AFM) was developed. 1 Instead of measuring ...
  21. [21]
    Scanning Probe Microscopy - Engineering and Technology History ...
    May 19, 2015 · They were invented in the early 1980s for basic research and quality control for microelectronics manufacturing, and they are still used in ...
  22. [22]
    Frictional Forces and Amontons' Law: From the Molecular to the ...
    This law states that for any two materials the (lateral) friction force is directly proportional to the (normal) applied load, with a constant of ...
  23. [23]
    Understanding the friction laws of Amontons and Coulomb by ...
    Dec 9, 2024 · We compared these results with the historical friction models of Amontons and Coulomb to better understand the actual differences between them ...
  24. [24]
    [PDF] The research works of Coulomb and Amontons and generalized ...
    Abstract: The paper is devoted to the contributions of Coulomb and Amontons to the physics of friction from the viewpoint of current discussions and attempts to ...
  25. [25]
    Friction - HyperPhysics Concepts
    When two surfaces are moving with respect to one another, the frictional resistance is almost constant over a wide range of low speeds, and in the standard ...Missing: definition | Show results with:definition
  26. [26]
    [PDF] Lecture 5 – Applying Newton's Laws - General Physics
    There are two kinds of friction: – Static friction: the maximum horizontal force that can be applied before an object starts to move. – Kinetic friction ...
  27. [27]
    [PDF] Static and Kinetic Friction
    Static friction is what keeps a resting body at rest. Kinetic friction is what slows down an object when slid on a surface.
  28. [28]
    Pesky squeaks and squeals caused by 3 types of 'stick-slip' behavior
    May 10, 2016 · Stick-slip is a frictional instability that governs diverse phenomena from squealing automobile brakes to earthquakes. At soft adhesive ...Missing: phenomenon | Show results with:phenomenon
  29. [29]
    What is the pitch of squeaking chalk? - MIT
    Once the block sticks back on the platform, the cycle repeats, causing the stick-slip phenomenon. Although the stick-slip behavior will cause an oscillation ...
  30. [30]
    Friction - Richard Fitzpatrick
    Friction is a force that impedes motion when a body slides over a rough surface. It is proportional to the normal reaction, and is calculated as f = μR_n.
  31. [31]
    [PDF] Kinetic Friction and Coefficients of Friction Introduction
    The empirical model for contact friction says the maximum friction force Ff = μN, where μ is the coefficient of friction (between the pair of surfaces) and N is ...
  32. [32]
    Deriving the dry friction law from fundamental laws of physics
    Jun 25, 2019 · The model shows that the dry friction law is consistent with the fundamental laws of physics, such as the dependencies between thermodynamic forces and flows.
  33. [33]
    12 Characteristics of Force - Feynman Lectures - Caltech
    There is another kind of friction, called dry friction or sliding friction, which occurs when one solid body slides on another. In this case a force is ...Missing: non- electromagnetic
  34. [34]
    Friction - Coefficients for Common Materials and Surfaces
    Rubber, Wet Concrete, Clean and Dry, 0.45 - 0.75. Silk, Silk, Clean, 0.25. Silver ... Typically steel on steel dry static friction coefficient 0.8 drops to 0.4 ...
  35. [35]
    EML2322L -- Friction Coefficients
    ### Summary of Friction Coefficients
  36. [36]
    [PDF] Coefficients of Friction Theory and Procedure
    Methods for Determining Coefficients of Friction. Force Probe Method. When dragging an object against a surface with a force probe at a constant velocity, the ...
  37. [37]
    Measuring Coefficient of Static Friction - Physics
    Measuring μs. An easy way to measure the coefficient of static friction is to place two objects together and then tilt them until the top one slides.
  38. [38]
    9.1 Dry Friction - Engineering Statics
    (9.1.1) 🔗 The friction angle is defined as the angle between the friction resultant and the normal force. At impending motion, the friction angle reaches its ...
  39. [39]
    Friction Angle - an overview | ScienceDirect Topics
    ... friction angle ϕp is: tan ϕ P = T → / N → = tan ϕ μ. T → and N → are the tangent and normal forces exerted on the surface by the solid when it starts moving.
  40. [40]
    Friction | Engineering Mechanics Review at MATHalino
    μ = Coefficient of friction. R = Resultant of f and N ϕ = angle of friction. Dry friction between the block and the floor. Formulas for dry friction. f=μN. tanϕ ...Problem 506 · Problem 509 · Problem 507 · Problem 508
  41. [41]
    experimental determination of the friction coefficient using a tilted ...
    Apr 11, 2015 · The objective of this article is show an easy way to measuring both dynamic and static friction factor between two solid surfaces, using Tilted plane method.
  42. [42]
    The Versatile Tilting-Plane Method of Measuring Coefficients of ...
    Sep 1, 2024 · This paper presents an improved tilting-plane method, which can measure these coefficients, i.e., µs, µk, and CR, for both sliding and rolling ...
  43. [43]
    33.4: Angle of Repose - Engineering LibreTexts
    Aug 24, 2023 · The angle of repose is the angle between the horizontal surface and the sloping surface of a conical pile of granular matter.
  44. [44]
    Angle of repose - DoITPoMS
    The angle of repose is the angle between the horizontal surface and the sloping surface of the pile. The tangent of this angle is the slope of repose.
  45. [45]
    A review on the angle of repose of granular materials - ScienceDirect
    May 1, 2018 · In soil mechanics, Karl Terzaghi defined the angle of repose as a special internal friction angle that is acquired under extreme (loosest state) ...
  46. [46]
    Mohr-Coulomb failure criteria - Geological Digressions
    Apr 3, 2024 · Tan φ is dimensionless. The internal friction angle is defined as the angle between the normal stress and resultant stress at the point of ...Missing: ≈ | Show results with:≈
  47. [47]
    Shear Strength of Soil and Mohr-Coulomb Theory
    Nov 10, 2024 · Φ is related to the internal friction that is predominant in soils like gravel and sand. This is why these soils are sometimes referred to as 'Φ ...Missing: geology ≈ 30°
  48. [48]
    [PDF] Slope Stability - USACE Publications
    Oct 31, 2003 · φ = angle of internal friction for Mohr-Coulomb diagram plotted in terms of total normal stress, σ (degrees) φ' = angle of internal friction ...
  49. [49]
    (a) Angle of friction (b) Angle of Repose (c) Cone of Friction
    Oct 22, 2019 · This inverted cone with semi-central angle, equal to limiting frictional angle θ, is called cone of friction. It is defined as the right ...
  50. [50]
    Angle and Cone of friction - Mechanics - CodeCogs
    Sep 11, 2011 · The relationship between the Angle and Cone of Friction - References for Angle and Cone of friction with worked examples.Missing: diagram | Show results with:diagram
  51. [51]
    The "ploughing" contribution to friction - IOPscience
    When a hard surface is slid over a softer surface part of the frictional resistance is due to the force required to plough asperities of the harder surface ...
  52. [52]
    8.6: Drag Forces in Fluids - Physics LibreTexts
    Jul 20, 2022 · The coefficient of viscosity η has SI units of [ N ⋅ m − 2 ⋅ s ] = [ Pa ⋅ s ] = [ kg ⋅ m − 1 ⋅ s − 1 ] ; a cgs unit called the poise is often ...
  53. [53]
    [PDF] CHAPTER 3 THE CONCEPT OF VISCOSITY
    Equation (3.1.3) is called the Newton's law of viscosity and states that the shear stress between adjacent fluid layers is proportional to the negative value of ...<|separator|>
  54. [54]
    Laminar and Turbulent Flow | Engineering Library
    All fluid flow is classified into one of two broad categories or regimes. These two flow regimes are laminar flow and turbulent flow.
  55. [55]
    [PDF] derivation of the stokes drag formula
    Adding things together we arrive at the famous Stokes Drag Law first derived by him over 150 years ago! Note that in this problem two thirds of the force on the.
  56. [56]
    Boundary Layer
    ... skin friction drag of an object, the heat transfer that occurs in high speed ... coefficient (mu). Re = V * r * l / mu. Boundary layers may be either ...
  57. [57]
    The Drag Equation
    The drag equation states that drag D is equal to the drag coefficient Cd times the density r times half of the velocity V squared times the reference area A.
  58. [58]
    Reynolds equation wiki - Tribonet
    Reynolds equation is a partial differential equation that describes the flow of a thin lubricant film between two surfaces.What is Reynolds equation? · Reynolds Number · Derivation of Reynolds Equation
  59. [59]
    Elastohydrodynamic Lubrication: Theory, Types & Practical Guide
    Elastohydrodynamic lubrication (EHL) describes a lubrication regime where high pressure causes significant elastic deformation of the contacting surfaces,
  60. [60]
    The Mechanism of Friction in Boundary Lubrication | J. Tribol.
    Scanning electron micrographs of the worn surfaces and surface profiles have shown that plowing is the dominant mechanism of friction in boundary lubrication.Missing: sources | Show results with:sources
  61. [61]
    Boundary Lubrication | About Tribology - Tribonet
    Mar 15, 2020 · Boundary lubrication describes situations where the applied load is primarily supported by contacting solid asperities, rather than a liquid lubricant.
  62. [62]
    Stribeck Curve | About Tribology
    ### Summary of Stribeck Curve Content
  63. [63]
    Mixed Lubrication Regime | About Tribology - Tribonet
    Aug 25, 2022 · The main purpose of the lubricant is to reduce the friction at the interface of the surfaces in contact. In the case of the mixed lubrication ...Background · Definition · Friction in mixed lubrication · Mixed lubrication models
  64. [64]
    A mixed lubrication model with consideration of starvation and ...
    Oct 11, 2012 · This article presents a mixed lubrication model for point contacts with considering the inlet starvation and interasperity cavitations.
  65. [65]
    The Hydrodynamic Study of the Swimming Gliding: a Two ... - NIH
    Oct 4, 2011 · Table 1. Drag coefficient values and contribution of pressure and friction drag for the total drag to each speed and for the four different ...
  66. [66]
    Skin-friction drag analysis from the forced convection modeling in ...
    This study deals with skin-friction drag analysis in underwater swimming. Although lower than profile drag, skin-friction drag remains significant and is ...Missing: coefficient | Show results with:coefficient
  67. [67]
    <i>In vivo</i> friction properties of human skin - O&P Library
    The average coefficient of friction for all measurements is 0.46±0.15 (p<0.05). Among all measured sites, the palm of the hand has the highest coefficient of ...
  68. [68]
    Introduction to Dynamic Mechanical Analysis and its Application to ...
    Tan δ indicates the relative degree of energy dissipation or damping of the material. For example, a material with a tan δ > 1 will exhibit more damping than a ...Missing: internal friction
  69. [69]
    [PDF] 10 Viscoelasticity
    viscoelastic rheological model will usually be of the form of the generalized Maxwell ... The constitutive equation for the Maxwell model is given by Eqn. 10.3.6,.
  70. [70]
    Proteoglycan 4 reduces friction more than other synovial fluid ...
    PRG4 in the presence of HA was found to significantly reduce the coefficient of friction for both cartilage-cartilage and cartilage-CoCrMo interface.
  71. [71]
    Rolling Resistance - The Engineering ToolBox
    Rolling Friction Coefficients ; 0.001 - 0.002 · 0.001 · 0.002 - 0.005 ; 0.5 ; railroad steel wheels on steel rails, ball bearings · bicycle tire on wooden track · low ...Missing: ice | Show results with:ice<|separator|>
  72. [72]
    Rolling Resistance Coefficient - an overview | ScienceDirect Topics
    Calculation of b is done through rolling resistance coefficient: b = R . C r . For ball-raceway contact, typical value of C r is 0.0015. Forces balance on ...
  73. [73]
    Rolling Resistance - an overview | ScienceDirect Topics
    Rolling resistance refers to the energy loss occurring in the tyre due to the deformation of the contact area and the damping properties of the rubber [79]. The ...
  74. [74]
    Modeling and Verification of Rolling Resistance Torque of High ...
    Mar 25, 2023 · Due to the viscoelasticity of rubber materials, hysteresis loss due to deformation is the main reason for the rolling resistance of high-speed ...
  75. [75]
    mechanics - rollingresistance
    Rolling Resistance When a wheel rolls on a surface energy is dissipated. Several sources of energy loss are active: i) work due to the flexing of the tire ...<|separator|>
  76. [76]
    [PDF] A literature study of rolling resistance and its affecting factors - kth .diva
    Jun 29, 2022 · Smaller deformations of the tire result in less energy loss due to hysteresis during tire deformation, hence decreasing rolling resistance [47].
  77. [77]
    [PDF] rolling resistance of pneumatic tires - ROSA P
    Due to mixing, the air inside the tire cavity will have a higher average temperature than at the be- ginning of the rolling process.
  78. [78]
    [PDF] Lecture Notes 14 - High Energy Physics
    ⇐ Abraham-Lorentz formula. This relation is known as the Abraham-Lorentz formula for the EM “radiation reaction” force. ( ). ( ). 2 r r. 6 o rad q. F t. a t c μ.
  79. [79]
    [PDF] arXiv:1204.5699v1 [quant-ph] 25 Apr 2012
    Apr 25, 2012 · The earliest work on radiation reaction is that of Abraham [1] and Lorentz [2], whose result is summarized in the well-known equation: Mx − Mτe.
  80. [80]
    [PDF] Self-Force and Radiation Reaction - UCSB Physics
    Self-force is the infinite force a particle exerts on itself. Radiation reaction force, from Maxwell equations, has pathological consequences. Point charges ...
  81. [81]
    Quantifying the triboelectric series | Nature Communications
    Mar 29, 2019 · The triboelectric effect is a type of contact-induced electrification, owing to which a material would become electrically charged after it ...
  82. [82]
    Quantifying and understanding the triboelectric series of inorganic ...
    Apr 29, 2020 · The triboelectric series describes materials' tendency to generate triboelectric charges. The currently existing forms of triboelectric series ...
  83. [83]
    [PDF] Triboelectric charge transfer theory driven by interfacial ... - arXiv
    Historically, triboelectricity has been characterized using a triboelectric series, which empirically indicates the positive and negative charge between two.
  84. [84]
    Minimizing friction, wear, and energy losses by eliminating contact ...
    Nov 16, 2018 · (A) Tribocharging causes high electrostatic adhesion, leading to high friction, wear, and energy consumption, as imaged in the temperature ...
  85. [85]
    Friction coefficient dependence on electrostatic tribocharging - Nature
    Aug 12, 2013 · To conclude, tribocharges produced by friction have a large effect on the friction coefficients of dielectrics that may exceed all other factors ...
  86. [86]
    [PDF] The mechanics of belt friction revisited - UC San Diego
    The increment of the force in the belt is then dT = μTdϕ, and the integration gives T2 = T1exp(μθ).Missing: source | Show results with:source
  87. [87]
    Nonlinear Dynamics Analysis of Disc Brake Frictional Vibration - MDPI
    Nov 26, 2022 · The braking force of the disc brake system comes from the friction between the brake disc and the brake pad. It is due to the nonlinearity of ...Missing: runaway | Show results with:runaway
  88. [88]
    [PDF] Rolling Bearings Handbook
    Rolling bearings use balls or rollers to reduce friction, with a coefficient of friction less than 1/100 of sliding bearings.
  89. [89]
  90. [90]
    [PDF] Ball Roller Bearings
    Friction coefficient l. Deep groove ball bearing 0.001 0 − 0.001 5. Angular contact ball bearing 0.001 2 − 0.002 0. Self-aligning ball bearing. 0.000 8 − 0.001 ...
  91. [91]
    [PDF] Air Bearing Design and Application Guide - IBS Precision Engineering
    Jun 7, 2023 · Being fluid film bearings, air bearings have a squeeze film damping effect resulting in higher dynamic stiffness and stability. No Lubrication.
  92. [92]
    Zero Friction - New Way Air Bearings
    Oct 3, 2012 · New Way Porous Media™ air bearings ... The result is that you then have almost infinite resolution, and the repeatability so critical to today's high performance, ...
  93. [93]
    The basics of air bearings - Linear Motion Tips
    The fluid film in an air bearing serves to average out small-scale errors, allowing air bearings to provide more accurate motion than mechanical bearings.Missing: zero | Show results with:zero
  94. [94]
    A prototype of an energy-efficient MAGLEV train - ScienceDirect.com
    This assembly prevents the vehicle from any direct mechanical contact to the track, thus avoiding any friction losses. ... This arrangement allows the vehicle and ...
  95. [95]
    How Maglev Works | Department of Energy
    Jun 14, 2016 · In Maglev, superconducting magnets suspend a train car above a U-shaped concrete guideway. Like ordinary magnets, these magnets repel one another when matching ...Missing: eliminates mechanical
  96. [96]
    Enhancing Tapered Roller Bearing Efficiency: The Impact of ...
    Improved Load Distribution: The even distribution of load in the rolling elements minimizes peak stresses and extends bearing life. Reduced Friction and Wear: ...
  97. [97]
    Optimization of partially crowned roller profiles for tapered roller ...
    Feb 19, 2017 · Two profile parameters of rollers employed for TRBs including the central flat length and crown radius were investigated. The optimal design ...Missing: gears | Show results with:gears
  98. [98]
    Active Vibration Control - an overview | ScienceDirect Topics
    Active vibration control refers to a system that employs mechanical and electronic components to detect and counteract unwanted vibrations using sensors, ...
  99. [99]
    [PDF] A new concept of active vibration controller through a fric
    Thus in this paper we will talk about active control for friction dampers. The active control is assured by the clamping force of a screw. The role of the ...
  100. [100]
    Active Damping Devices - Micromega Dynamics
    In structural active damping, sensors are used to detect vibrations or resonances in a structure. These sensors send signals to an active damping system ...
  101. [101]
    [PDF] Boundary Lubrication--Revisited
    The shear strength of a boundary lubricating film should be directly reflected in the friction coefficient. In general, this is true with low shear strength ...
  102. [102]
    [PDF] Grease Types And Grades
    Base Oil: Usually mineral oil, synthetic oil, or vegetable oil that provides the lubrication properties. Thickener: A soap or non-soap based substance that ...
  103. [103]
    [PDF] Synthetic versus Mineral Fluids - in Lubrication
    Synthetic lubricants may offer improved energy efficiency, wider temperature range, and better reliability, and can be more expensive initially. Mineral oils ...
  104. [104]
    [PDF] Examining Interactions Between Lubricant Components - OSTI.GOV
    ... MoS2 or similar, while the most common antiwear additive in lubricants ... lubricant components such as friction modifier Molyvan L and anti-wear additive ZDDP.Missing: disulfide | Show results with:disulfide
  105. [105]
    Modelling Friction Reduction Based on Molybdenum Disulphide ...
    Mar 17, 2025 · Molybdenum dialkyldithiocarbamate is a highly effective friction modifier lubricant additive in boundary lubrication, owing to the formation ...Missing: anti- | Show results with:anti-
  106. [106]
    Understanding the Friction Reduction Mechanism Based ... - PubMed
    Nov 13, 2018 · Among friction modifier lubricant additives, molybdenum dialkyldithiocarbamate (MoDTC) provides excellent friction behavior in boundary ...Missing: anti- | Show results with:anti-
  107. [107]
    [PDF] Selection of Motor Oil - Open PRAIRIE - South Dakota State University
    Viscosity index of an oil indicates amount of viscosity change in a temperature change from 0° to. 210°F. Low viscosity index means a large change and high ...
  108. [108]
    [PDF] Relationships between physical properties and chemical ...
    In addi- tion to Mikeska's [1] values, the viscosity and viscosity-index charts contain values for 1,I-dicyclohexylhexadecane and l-cyclohexyl-2 ...<|control11|><|separator|>
  109. [109]
    [PDF] PRE-PUBLICATION NOTICE - EPA
    Sep 20, 2024 · Environmentally acceptable lubricant (EAL) means a lubricant or hydraulic fluid, including any oil or grease, that is “biodegradable ...
  110. [110]
    Friction, Important Attribute of Tires, Roads and Deicers
    Oct 2, 2018 · A patterned tire gives typical dry and wet frictional coefficients of about 0.7 and 0.4, respectively. These values represent a compromise ...
  111. [111]
    Developments in tyre design for lower rolling resistance: A state of ...
    Many studies reveal that the tyre's rolling-resistance can be responsible for up to approximately 20–30% of the vehicle's total fuel consumption depending on ...
  112. [112]
    Brake Disc vs. Drum Brakes: Which Is Right for You? - MAT Foundry
    Apr 11, 2024 · Drum Brakes: While drum brakes might not excel in performance compared to disc brakes, they are still effective for everyday driving conditions.
  113. [113]
    Quantifying the change of brake wear particulate matter emissions ...
    Nov 1, 2023 · The results show that RBS would reduce brake wear by between 64 and 95%. The study highlights the effect of aggressive braking on the amount of friction brake ...
  114. [114]
    Friction and Ancillary Losses - DieselNet
    Engine mechanical losses result from friction in the power cylinder unit, crankshaft, and valvetrain, as well as from ancillary loads such as pumps and fans.
  115. [115]
    [PDF] Friction Losses between Piston Ring-Liner Assembly of Internal ...
    Abstract- In IC engine piston ring friction losses account for approximately 20% of total mechanical losses as reported in the literature.
  116. [116]
    [PDF] Evaluation of Skin Friction Drag for Liner Applications in Aircraft
    Drag is recognized as a significant issue in aircraft design with one estimate by Malik, et al. 1 that skin friction constitutes approximately 50% of the drag ...
  117. [117]
    Tribological coatings for electric vehicle applications - Frontiers
    These coatings may also work equally well in EVs by providing low friction and wear combined with a high dielectric constant or insulating properties to ...
  118. [118]
    Global energy consumption due to friction in trucks and buses
    Aug 6, 2025 · Friction is roughly responsible for 23 % of the global energy consumption and approximately 33 % of the energy used in transportation (Holmberg ...
  119. [119]
    Coefficient of Friction Testing: Procedure, Application, Benefits, and ...
    Feb 13, 2024 · Another arrangement for testing the coefficient of friction is the inclined plane method. ... These include pin-on-disk/ball-on-disk tribometer ...
  120. [120]
  121. [121]
    G99 Standard Test Method for Wear and Friction Testing with a Pin ...
    Nov 14, 2023 · This use should be described as having “followed the procedure of ASTM G99.” All test parameters that were used in such case must be stated.
  122. [122]
    Pinning Down the Pin-on Disk Tribometer
    The Pin-on Disk Tribometer is a test instrument designed for precise and repeatable tribological characterization of bulk materials, coatings and lubricants.Missing: inclined | Show results with:inclined
  123. [123]
    [PDF] Wear Testing with a Pin-on-Disk Apparatus1 - ResearchGate
    Jan 1, 2017 · This use should be described as having “followed the procedure of ASTM G99.” All test parameters that were used in such case must be stated. 1.4 ...
  124. [124]
    Brake Inertia Dynamometer Test for Corrosion Effects on Friction ...
    30-day returnsThe current inertia dynamometer tests published by Standard Development Organizations assess friction behavior, wear, noise, vibration, and harshness under ...
  125. [125]
    Friction Determination by Atomic Force Microscopy in Field of ... - NIH
    Jun 21, 2018 · This paper will introduce the principles of mechanical measurement by using AFM and reviewing the progress of AFM methods in determining frictions in the field ...
  126. [126]
    Optical Measurement of Oil Film Thickness under Elasto ... - Nature
    The use of Newton rings for the measurement of oil film thickness in point or line contact is rendered difficult by the similarity of the refractive index of ...
  127. [127]
    [PDF] OPTICAL INTERFEROMETRY STUDY OF FILM FORMATION IN ...
    Using white-light interferometry it was possible to observe directly the thickness of the lubricating film in the contact under conditions of pure rolling ...
  128. [128]
    [PDF] Uncertainty Analysis for Friction Coefficient Measurements
    Laboratory experimentation remains the only practical method available for the accurate identification of friction coefficients for arbitrary material pairs.
  129. [129]
    Contaminant film thickness affects walkway friction measurements
    Aug 29, 2022 · Friction measured with the Mark IIIB decreased with increasing film thickness on one surface across all three contaminants and on a second ...
  130. [130]
    Research on Measurement Methods of Friction Coefficient
    ... accuracy of up to 0.01s. 2.2 Measurement method based on the plane method. 2.2.1 Electrocontrolled translation and static friction coefficient meter. Peng ...
  131. [131]
    Tribological characteristics of graded pencil cores on paper
    Static friction coefficients were determined from simple, inclined-plane tests of the various core-paper combinations, and kinetic friction coefficients were ...
  132. [132]
    The Chemistry of Matches - Compound Interest
    Nov 20, 2014 · When the match is struck, a small amount of the red phosphorus on the striking surface is converted into white phosphorus, which then ignites.
  133. [133]
    (PDF) Enhancement of Deep Drilling for Stainless Steels by Nano ...
    Oct 14, 2025 · This paper represents a new lubricant method which is able to one-stroke drill deep holes with a length-to-diameter of 8, on the AISI SUS 304 ...
  134. [134]
    [PDF] Variant Friction Coefficients of Lagging and Implications for ...
    Designers use the friction coefficient in the Euler Capstan equation to calculate the drive capacity of the conveyor, so the behavior of lagging friction under ...
  135. [135]
  136. [136]
    Global energy consumption due to friction and wear in the mining ...
    Friction and wear is annually resulting in 970 million tonnes of CO2 emissions worldwide in mineral mining (accounting for 2.7% of world CO2 emissions).
  137. [137]