Next Article in Journal
Combined Geophysical–Geodynamic Analysis of the Plio-Pleistocene Hominin Dispersal Through the Levantine Corridor
Previous Article in Journal
Systematic Monte Carlo Analysis of Binary Compounds for Neutron Shielding in a Compact Nuclear Fusion Reactor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Comprehensive Review of Rubber Contact Mechanics and Friction Theories

by
Raffaele Stefanelli
1,*,
Gabriele Fichera
2,*,
Andrea Genovese
1,
Guido Napolitano Dell’Annunziata
1,
Aleksandr Sakhnevych
1,
Francesco Timpone
1 and
Flavio Farroni
1
1
Department of Industrial Engineering, University of Naples Federico II, 80125 Naples, Italy
2
Department of Civil Engineering and Architecture, University of Catania, 95123 Catania, Italy
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2025, 15(21), 11558; https://doi.org/10.3390/app152111558
Submission received: 2 October 2025 / Revised: 21 October 2025 / Accepted: 28 October 2025 / Published: 29 October 2025

Abstract

This review surveys theoretical frameworks developed to describe rubber contact and friction on rough surfaces, with a particular focus on tire–road interaction. It begins with classical continuum approaches, which provide valuable foundations but show limitations when applied to viscoelastic materials and multiscale roughness. More recent formulations are then examined, including the Klüppel–Heinrich model, which couples fractal surface descriptions with viscoelastic dissipation, and Persson’s theory, which applies a statistical mechanics perspective and later integrates flash temperature effects. Grosch’s pioneering experimental work is also revisited as a key empirical reference linking friction, velocity, and temperature. A comparative discussion highlights the ability of these models to capture scale-dependent contact and energy dissipation while also noting practical challenges such as calibration requirements, parameter sensitivity, and computational costs. Persistent issues include the definition of cutoff criteria for roughness spectra, the treatment of adhesion under realistic operating conditions, and the translation of detailed power spectral density (PSD) data into usable inputs for predictive models. The review emphasizes progress in connecting material rheology, surface characterization, and operating conditions but also underscores the gap between theoretical predictions and real tire–road performance. Bridging this gap will require hybrid approaches that combine physics-based and data-driven methods, supported by advances in surface metrology, in situ friction measurements, and machine learning. Overall, the paper provides a critical synthesis of current models and outlines future directions toward more predictive and application-oriented tire–road friction modeling.

1. Introduction

Tires are the key functional link between a vehicle and the road, transmitting the forces that enable acceleration, braking, steering, and stability. Their performance directly affects handling, comfort, safety, and energy efficiency. At the core, tire behavior is governed by the interaction between rubber and the rough road surface, which determines the available grip. Understanding this interaction requires a solid foundation in contact mechanics and friction theory [1,2,3,4,5].
Beyond tire–road applications, closely related rubber–shell interactions occur in frictional dampers, shell-type shock absorbers, and elastomeric elements radially confined within curved housings. In these systems, contact mechanics controls local conformability and energy dissipation under confinement/curvature. Both analytical models (e.g., interaction of elastomeric fillers with open shells; structural damping in shell shock-absorber friction modules) and numerical studies (e.g., simulation of novel frictional dampers; adaptive behavior of shell-type elastic elements for drilling shock absorbers) show how viscoelastic contact principles extend to curved geometries [6,7,8,9].
Contact mechanics, a core discipline of tribology, investigates how solid bodies interact when in contact, whether under static or dynamic conditions [10,11,12,13,14,15]. Rooted in materials science and continuum mechanics, it provides a framework to analyze deformation under load, the distribution of stresses and pressures at interfaces, and the emergence of frictional forces [16,17,18,19]. High local stresses often develop at contact points, which may lead to fracture, yielding, fatigue, or wear. A key concept is therefore the determination of contact pressure, strongly influenced by geometry, material properties, and applied load [20,21,22,23]. Adhesion also plays a significant role, as intermolecular attraction between surfaces contributes to static friction, energy dissipation, and wear, particularly in soft materials such as rubber [24,25,26,27].
In practice, tire–road interfaces involve surfaces with pronounced roughness and curvature, resulting in highly localized contacts. The actual contact area is much smaller than the nominal footprint, generating elevated stresses and complicating friction prediction. For rubber, the degree of penetration into road asperities is critical, as it controls both hysteretic energy loss and adhesion [28,29,30]. This effect is commonly quantified by the ratio of real to nominal contact area ( A c / A 0 ), where A 0 is the projected footprint and A c is the true contact area. This ratio links surface roughness, rubber viscoelasticity, and loading conditions and is widely used in predictive friction models.
Ultimately, an accurate description of the A c / A 0 ratio is central to understanding interfacial phenomena such as thermal and electrical resistance, adhesion, wear, and frictional behavior [31,32]. Modern contact theories therefore integrate realistic topographies, viscoelastic material response, and load dependence to provide a more reliable description of tire–road interaction.

2. Overview of Contact Mechanics

Classical contact mechanics established the analytical foundations for describing stresses, deformations, and real contact areas under load. Although developed for smooth, elastic bodies, these models remain the starting point for modern extensions to rough and viscoelastic surfaces. The most influential formulations include Hertz’s solution for elastic contact, statistical models that incorporate surface roughness, and adhesive theories that account for intermolecular forces.

2.1. Hertzian Contact Theory

The foundation of modern contact mechanics can be traced to Heinrich Hertz’s seminal work in 1882 [33]. Hertzian theory describes the localized stresses and deformations that arise when two elastic, axisymmetric, parabolic bodies are pressed together under a normal load [33,34,35]. It provides closed-form expressions for contact area, pressure distribution, and surface displacements as a function of applied force, radii of curvature, and elastic properties. Importantly, the solution applies to non-adhesive contacts, where no tensile stresses occur in the contact patch. Despite its assumptions, Hertzian theory remains a cornerstone for more advanced models.
Although originally derived for parabolic geometries, Hertzian solutions remain accurate when applied to practical cases such as sphere–sphere and cylinder–cylinder contact [36,37,38]. In these cases, theoretical point or line contacts are replaced by finite contact areas due to elastic deformation, leading to so-called Hertzian contact stresses [39,40]. The theory assumes small deformations, an unchanged geometry, and a purely elastic response. For two cylinders, the pressure distribution is nearly parabolic, peaking at the center, where the maximum contact pressure p m a x occurs, and vanishing at the edges.
A simplified yet widely used configuration is the indentation of an elastic half-space by a rigid sphere, illustrated in Figure 1.
In this case, the indentation depth d and the contact radius a can be expressed analytically in terms of the load F, sphere radius R, and effective elastic modulus E * .
The principal Hertzian equation for a spherical elastic body indenting on an elastic half-space is reported in [20,26]. Assuming a quadratic pressure distribution:
p ( r ) = p m a x 1 r 2 a 2 , r a ,
the resulting force–displacement relationship takes the form
F = 4 3 E * R 1 / 2 d 3 / 2 ,
from which the key Hertzian expressions follow:
a = 3 F R 4 E * 1 / 3 , p m a x = 6 F E * 2 π 3 R 2 1 / 3 .
The effective elastic modulus E * accounts for the properties of both bodies:
1 E * = 1 ν 1 2 E 1 + 1 ν 2 2 E 2 ,
where E i and ν i are Young’s modulus and Poisson’s ratio of body i.
Figure 2 further illustrates the quadratic pressure distribution under increasing vertical loads, showing how the maximum pressure grows while the contact radius expands accordingly.
Although Hertz and subsequent extensions by Sneddon [41,42] did not account for adhesion, this framework remains the analytical basis for many modern models of rough surface contact, where asperities are idealized as spherical indenters on an elastic half-space.

2.2. Surface Roughness in Contact Mechanics

Incorporating surface roughness into contact models presents a major challenge, given the stochastic and multiscale nature of real topographies [43,44,45]. A common approach is to describe rough surfaces as ensembles of asperities, each treated individually using Hertzian contact mechanics [46].
The pioneering work of Bowden and Tabor (1939) revealed that the real contact area between two rough surfaces is only a fraction of the nominal footprint [47]. Their observations demonstrated that asperity-level interactions dominate the stress distribution and can induce yielding or failure [16], establishing the basis of modern tribology [48,49,50,51].
Building on this paradigm, Archard [52,53,54] explained the experimentally observed linear relation between real contact area and applied load, which the classical Hertzian theory could not capture. He proposed a multi-asperity model where overall contact arises from the superposition of many Hertzian microcontacts, with the number of asperities scaling linearly with load. Archard also recognized the inherently multiscale nature of roughness [55,56], anticipating later fractal descriptions [57,58,59,60,61].
Following Archard, Greenwood and Williamson (1966) and Bush et al. (1975) provided rigorous mathematical frameworks for statistical modeling of rough surfaces [20,22,62,63]. The Greenwood–Williamson (GW) model is particularly influential, describing a rough surface as a population of spherical asperities with identical radii but randomly distributed summit heights. Each asperity deforms independently under Hertzian theory, and the overall contact is obtained by integrating the contributions of all asperities [64,65,66,67]. At its core, the GW model assumes that the asperities deform independently and that their summits are spaced sufficiently apart to neglect mutual interactions [68].
For rubberlike materials, the assumption of identical, non-interacting spherical asperities introduces some inaccuracies because the large compliance of the polymer causes neighboring contact spots to elastically interact through overlapping deformation fields. Finite-element and boundary-element analyses have shown that this simplification can overestimate the real contact area by 20–40% under typical tire–road pressures and underestimate the local peak stresses by a similar factor.
As illustrated in Figure 3 and Figure 4, the spherical asperities higher than the mean separation d contribute to the contact.
The contact radius a i , area A i , and load F i of a single asperity with height z are
a i = R ( z d ) , A i = π R ( z d ) , F i = 4 3 E * R 1 / 2 ( z d ) 3 / 2 ,
where E * is the effective elastic modulus (Equation (4)).
The total real contact area and load follow from integration over the asperity height distribution ϕ ( z ) :
A c = π N R d ( z d ) ϕ ( z ) d z , F = 4 3 E * N R 1 / 2 d ( z d ) 3 / 2 ϕ ( z ) d z ,
where N is the total number of asperities. Introducing the surface density η with N = η A 0 , Greenwood and Williamson expressed these integrals in terms of the normalized separation t = d / σ and the dimensionless functions F n ( t ) :
F n ( t ) = t ( z t ) n ϕ * ( z ) d z ,
where ϕ * ( z ) is the standardized distribution [20,62].
The GW model thus links microscopic roughness features to macroscopic contact parameters such as load and real contact area, providing a powerful statistical framework that remains widely used in analytical and computational tribology [70,71,72,73].
Bush et al. [63,74] extended the GW formulation by adopting paraboloidal asperity summits and confirmed the linear scaling of real contact area with load under small deformations. Despite the simplification of neglecting asperity interactions, both the GW and Bush models remain indispensable in tribology, especially for applications involving rubber–road interaction and tire dynamics.

2.3. The Role of Adhesion in Contact Mechanics

Adhesion plays a crucial role in contact mechanics, especially for soft materials such as elastomers [75,76,77,78]. Attractive intermolecular forces between solids in close proximity explain why a finite load is needed to separate two bodies in intimate contact [79,80,81]. This effect is particularly relevant under light loads [24,82]. Experimental studies confirmed that contact areas at low loads are significantly larger than Hertzian predictions and remain finite as the load approaches zero [83,84], consistent with earlier estimates by Bradley [85].
Compared to Hertz, the JKR theory provides a more accurate description of soft materials, as adhesion enlarges the real contact area and thus the friction coefficient, leading to closer agreement with experimental evidence (Figure 5).
Hertzian theory, while foundational, neglects adhesion [33]. Real surfaces, however, experience adhesive interactions at the molecular level that enlarge the contact area beyond purely elastic predictions [25,87,88]. To address this, two classical adhesive contact models were introduced: the Johnson–Kendall–Roberts (JKR) theory [89] and the Derjaguin–Muller–Toporov (DMT) theory [90]. Both extend Hertz’s formulation by including van der Waals forces.
The JKR model accounts for adhesion within the contact region, predicting a finite contact radius even at zero load [91,92,93,94]. By minimizing the total energy, Johnson et al. derived the contact radius as
a 3 = 3 R 4 E * F + 3 π R γ + 6 π R F γ + ( 3 π R γ ) 2 ,
where R is the reduced radius of curvature, E * is the effective elastic modulus (Equation (4)), F is the applied load, and γ is the surface energy. At γ = 0 , the Hertz solution is recovered. JKR also predicts a pull-off force:
F J K R = 3 π R γ 2 ,
and a pressure distribution modified by adhesion [89].
Although the original JKR formulation addressed the contact of a sphere on a plane, its methodology is extendable to other geometries, such as sinusoidal surfaces [95].
The DMT theory, in contrast, retains Hertzian stresses inside the contact area but adds adhesive forces acting outside the contact area, modeled via a Lennard–Jones potential [96,97,98,99]. Thus, JKR incorporates adhesion inside the contact area, whereas DMT applies it externally. The schematic comparison between JKR and DMT is reported in Figure 6.
Although conceptually different, both JKR and DMT reduce to the Hertz solution in the absence of adhesion [100]. Initially treated as competing models, their applicability was clarified by Tabor [101], Muller [102] and Maugis [103], who showed that JKR is appropriate for soft, highly adhesive materials, while DMT better describes stiff, weakly adhesive systems [104].
Maugis later introduced a unified formulation [105,106,107,108], parameterized by a dimensionless Maugis parameter λ , which quantifies the relative importance of adhesion versus elasticity:
  • λ 1 : DMT regime;
  • λ 1 : JKR regime.
This parameter corresponds physically to the ratio between adhesion-induced elastic deformation and the range of adhesion forces [109,110].
For typical tire tread compounds, the Maugis parameter λ generally falls within the intermediate transition regime, implying partial adhesion within the contact region. Experimental investigations on filled SBR and NR blends have shown that λ decreases with increasing temperature—mainly due to the reduction in surface energy γ and the elastic stiffening of the rubber matrix. Elevated humidity or hydrophilic surface contamination tends to lower the effective surface energy, shifting the contact behavior toward the DMT limit, whereas clean, dry interfaces reinforced with silica or carbon black exhibit higher λ values. Therefore, real tire–road interactions typically operate in a mixed or transitional regime where both adhesive and elastic contributions coexist.
In summary, adhesive contact models (JKR, DMT, and Maugis’ transition theory) demonstrate that adhesion introduces nonlinear relationships between load and contact area, highlighting the limitations of purely elastic theories and the critical role of surface energy in realistic contact mechanics.

2.4. Modern Theories of Contact Mechanics: Rubber Block Sliding on Rough Surfaces

Over the past few decades, significant advances in contact mechanics have been achieved through both experimental tools and numerical simulations [111,112,113,114]. A key contribution is the semi-analytical framework of Klüppel and Heinrich, who extended the classical Greenwood and Williamson (GW) model to describe the sliding of rubber on rough surfaces [115,116]. Their approach incorporates the interplay between the viscoelastic properties of rubber and the multi-scale nature of roughness, accounting for both adhesive and hysteretic friction contributions [117,118,119,120,121,122,123,124].
While the GW theory effectively models hard materials by assuming independent asperities [62], this assumption fails for rubber, where compliant deformation fields overlap at small scales. To address this, Klüppel and Heinrich adapted the GW model to include asperity interactions, thus enabling a more realistic description of the real contact area under sliding with velocity v and vertical load F (Figure 7).
In this formulation, the original height distribution ϕ ( z ) is replaced by the summit height distribution ϕ s ( z ) , obtained through an affine transformation, which better describes the portion of the rough profile effectively in contact [116,125]. Introducing the standardized separation t s = d / σ s , the GW functions F n ( t s ) can be evaluated, with F 0 ( t s ) denoting the probability of summit–rubber contact [115,125].
The model further incorporates the viscoelastic nature of rubber through its dynamic complex modulus E * ( ω ) [126,127,128,129]. Under standard assumptions ( ν 1 = 0.5 , E 1 E 2 ), the effective modulus becomes [125]
1 E * = 1 ν 1 2 E 1 + 1 ν 2 2 E 2 3 4 | E * ( ω ) |
Considering the largest asperities as spherical caps of radius R, the normal load and stress can be expressed in terms of the GW functions and the summit height distribution as [115,125,130]
F = 16 9 A 0 N s | E * ( ξ | | ) | R 1 / 2 σ s 3 / 2 F 3 / 2 ( t s ) ,
σ 0 = F A 0 .
Similarly, the macroscopic real contact area A c , 0 is obtained as a fraction of the nominal area A 0 (Figure 8) and can be related to the fractal dimension D of the surface roughness [115,125,131]:
A c , 0 = ( 2 D 4 ) 12 3 ( 2 D 2 ) A 0 F 0 ( t ) .
This fractal characterization is often performed through statistical descriptors such as the Height Difference Correlation (HDC) function [132,133,134].
Because roughness is inherently multi-scale, the real contact area A c is scale-dependent. The yardstick method [115,135] illustrates this by showing how the apparent contact area increases as smaller cavities are accounted for:
A c ( λ ) = A c , 0 λ ξ | | 2 D , λ < ξ | | .
The smallest scale at which rubber conforms, λ m i n , is determined by an energy balance between deformation, adhesion, and stored elastic energy [115]. This defines the effective real contact area A c ( λ m i n ) , which is generally below 10% of the nominal contact area A 0 at high velocities and low loads [116].
In practical numerical implementations, λ min is obtained by numerically solving the energy-balance condition for the unknown wavelength under the assigned operating conditions (velocity, contact pressure, and temperature). The computation uses the viscoelastic modulus E * ( ω ) —evaluated at ω = 2 π v / λ via time–temperature superposition—and the surface spectral descriptors as inputs. Hence, λ min represents the smallest wavelength at which the deformation energy of rubber still compensates for the combined elastic and adhesive contributions.
Recognizing that road surfaces are better described by two scaling regimes (macro- and micro-texture), Klüppel and Heinrich extended their theory accordingly [115,125]. The contact area expression then becomes
A c ( λ ) = A c , 0 λ 2 ξ | | 2 D 1 λ m i n λ 2 2 D 2 .
The adoption of two scaling regimes is not arbitrary but reflects the two distinct slopes typically observed in asphalt spectra (macro- vs. micro-texture).
Some results from this theory, showing the ratio A c / A 0 as a function of sliding velocity and pressure, are reported in Figure 9. Increasing velocity reduces A c as rubber becomes glassy, while higher contact pressure increases indentation depth and enlarges the real area of contact. Overall, A c remains a small fraction of A 0 .
In summary, the Klüppel–Heinrich extension of the GW framework captures the scale-dependent behavior of rubber sliding on rough surfaces by combining fractal descriptors of roughness with the viscoelastic properties of elastomers. This provides a solid physical basis for the subsequent development of tire friction models.
While Klüppel and Heinrich’s framework provides a physically grounded extension of GW to viscoelastic rubber on multiscale roughness, several assumptions (e.g., discrete asperities, limited coupling across scales) motivate alternative descriptions. Persson’s theory instead treats surface roughness as a continuous random field and formulates contact as a scale-evolving stochastic process in stress space, offering a complementary and often more general perspective on rubber–road interactions [136,137,138,139,140,141].
In Persson’s approach, the interface is examined at increasing magnifications ζ = L / λ , where L is the macroscopic contact diameter and λ is the smallest resolved length scale [76,142]. As ζ increases, new roughness wavelengths become visible and previously apparent contact regions are revealed as (partly) non-contact regions. The relative contact area at magnification ζ is
P ( ζ ) = A ( ζ ) A 0 , with σ 0 A 0 = σ ζ A ( ζ ) ,
so that P ( ζ ) = σ 0 / σ ζ . The average local pressure σ ζ follows from the stress probability density P ( σ , ζ ) :
σ ζ = 0 σ P ( σ , ζ ) d σ 0 P ( σ , ζ ) d σ .
The evolution of P ( σ , ζ ) is governed by a diffusion-like equation in which “time’’ is replaced by magnification [142]:
P ( σ , ζ ) ζ = f ( ζ ) 2 P ( σ , ζ ) σ 2 ,
supplemented by (i) a reflecting/absorbing boundary at zero tension in the non-adhesive case [143], P ( 0 , ζ ) = 0 , and (ii) the initial condition P ( σ , ζ = 1 ) = δ ( σ σ 0 ) , i.e., uniform macroscopic pressure σ 0 . The diffusion coefficient couples elasticity and roughness via the power spectral density (PSD) C ( q ) :
f ( ζ ) = π 4 E 1 ν 2 2 q ( ζ ) 2 C q ( ζ ) , q ( ζ ) = q L ζ , q L = 2 π L ,
where E is the effective modulus and ν is the Poisson ratio [142,144].
Solving (18) yields the contact fraction at scale q:
P ( q ) = 2 π 0 sin x x exp x 2 G ( q ) d x ,
with the cumulative roughness–rheology functional
G ( q ) = 1 8 q L q q 3 C ( q ) 0 2 π E ( q v cos ϕ ) ( 1 ν 2 ) σ 0 2 d ϕ d q ,
where E ( ω ) is the complex (frequency-dependent) modulus evaluated at ω = q v cos ϕ , with v being the sliding speed and ϕ being the angle between q and the sliding direction [142].
For σ 0 | E ( ω ) | , the small-x limit of (20) gives the widely used approximation
P ( q ) π G ( q ) 1 / 2 ,
and an accurate interpolation valid across regimes is [142]
P ( q ) 1 + π G ( q ) 3 / 2 2 / 3 .
The real contact area at the finest resolved scale q 1 is therefore
A c = A 0 P ( q 1 ) ,
with lower and upper cutoffs ( q 0 , q 1 ) bounding the self-affine PSD (macro to micro/atomic scales). In practice, q 1 reflects physical limitations (contamination layers, surface films, molecular smoothing), ensuring a finite A c even if the mathematical self-affinity extends to very high q [142].
Persson proposed an empirical rule for q 1 based on the rms slope of the surface:
h rms 2 = 1 2 π q 0 q 1 C 2 D ( q ) d q 1.3 ,
which fixes the upper cutoff by imposing a target slope level [111,145]. The upper cutoff is thus defined following Persson’s slope-weighted description of roughness, which properly accounts for short-wavelength features rather than relying solely on height amplitudes; this prevents bias in the real contact area estimate by correctly weighting spatial frequencies.
While convenient, this method does not capture potential changes in rubber behavior and chemistry at very small scales [146,147], and, more broadly, the choice of ( λ 0 , λ 1 ) (or ( q 0 , q 1 ) ) remains debated due to the lack of predictive mechano-chemical models for surface evolution under local stress/temperature.
Unlike asperity-based models, Persson’s theory (i) directly embeds the multiscale PSD into the contact solution, (ii) naturally incorporates viscoelastic dispersion via E ( ω ) at each q-scale, and (iii) predicts that P decreases with increasing magnification, while friction can remain finite because energy dissipation accumulates across scales [136,137,139]. This statistical mechanics view has become a cornerstone for modern friction modeling and will underpin later developments discussed in the next section.
Beyond Persson, recent work by Müser and co-authors has analyzed nonlinearity and nonlocality in soft–hard contact, elucidating how microscale morphology and long-range elastic coupling regulate contact area and interfacial forces [148,149].

3. Fundamentals of Friction Theories

While contact mechanics establishes a framework for stresses and deformations at interfaces, friction theories extend this foundation to explain how these interactions translate into energy dissipation and grip. Rubber friction is a complex phenomenon influenced by road roughness, contact area geometry, compound viscoelasticity, sliding velocity, and interface temperature. Since the forces exchanged between the tire and road are fundamentally frictional, understanding these mechanisms is crucial for tire performance analysis [150]. Over time, numerous theories have been developed to describe tire–road interaction [151,152,153]. The effective friction coefficient is generally attributed to three contributions: hysteresis, adhesion, and wear [154,155].
The hysteresis component (or indentation friction) originates from the internal viscoelastic deformation induced by asperities. As rubber slides, the time-dependent response generates asymmetric pressure fields that oppose motion [156,157,158,159].
The adhesive component arises from molecular interactions (e.g., van der Waals forces) forming at the true contact spots [160,161,162,163,164,165,166]. During sliding, bonds form, stretch and break, with viscoelastic resistance contributing to friction. At higher sliding velocities, reduced bonding time lowers the adhesive contribution [25,167,168]. Adhesion also increases the effective area involved in hysteretic losses [154].
The third contribution, wear-related friction, occurs when sharp asperities excite the rubber at high frequencies, possibly inducing local microfractures and energy dissipation [169]. However, this effect is usually negligible compared to adhesion and hysteresis and was already considered minor in practical conditions by the 1960s [154].
Based on these considerations, the total friction force F is expressed as [154]
F = F a + F h
where F a and F h are the adhesive and hysteretic contributions, respectively. A schematic illustration of these mechanisms is shown in Figure 10.
By dividing Equation (26) by the normal load, one obtains the total friction coefficient:
μ = μ a + μ h
where μ a and μ h denote the adhesive and hysteretic components. In wet conditions, the adhesive term is strongly reduced by the water film, so most of the grip arises from hysteresis.
Experimentally, the two mechanisms can be separated by repeating friction tests on the same rubber/roughness pair with and without a thin inert lubricant (or dusting), which strongly suppresses interfacial shear while leaving bulk viscoelastic hysteresis essentially unaffected, so that the difference estimates the adhesive share. Although such procedures cannot fully remove adhesion, they provide a practical means to quantify its relative contribution. From a computational standpoint, adhesion is mainly associated with the low-frequency range corresponding to slow interfacial slip, whereas hysteresis dominates the higher frequencies linked to bulk viscoelastic deformation. These distinct spectral domains—interfacial for adhesion and bulk for hysteresis—naturally prevent double counting when the two mechanisms are superposed in hybrid models.
Over the years, several experimental and theoretical contributions have advanced the understanding of these mechanisms, as will be discussed in the following sections.

3.1. Grosch’s Experimental Tests

One of the pioneers in rubber friction studies was Grosch, who conducted systematic experiments to identify the main mechanisms of friction [170,171,172,173]. He showed that rubber friction is primarily governed by adhesion and hysteresis and related these contributions to the viscoelastic properties of rubber compounds.
Grosch tested four rubbers (INR, ABR, SBR, butyl) against smooth glass and rough silicon carbide. Friction was measured using a rig where the sample was pressed against the surface under controlled temperature (−50 ∘C to 100 ∘C) and humidity, with sliding velocities up to 30 mm/s to avoid heating effects. The data were shifted into master curves using the Williams–Landel–Ferry (WLF) equation [174,175,176].
For SBR and ABR, two distinct peaks appeared in the master curves: a low-velocity peak associated with adhesion and a high-velocity peak attributed to hysteresis. Suppressing molecular contact by dusting the surface reduced only the first peak, confirming this interpretation.
Two characteristic length scales emerged from these results. For adhesion, Grosch found a constant ratio between the velocity at the first peak and the frequency of maximum loss modulus, corresponding to a molecular slip distance of about 6 × 10 9 m. For hysteresis, the velocity of the second peak was related to the frequency of maximum loss tangent, corresponding to a length scale of 1.5 × 10 4 m, consistent with the average asperity spacing.
These characteristic lengths should not be regarded as universal constants but as empirical correlations arising from the interplay between the viscoelastic spectrum of the compound and the surface morphology. Their magnitude can vary with filler morphology, crosslink density, temperature, and surface cleanliness, yet their order of magnitude remains consistent across typical tire rubbers and test surfaces, supporting Grosch’s mechanistic interpretation of adhesive and hysteretic friction processes.
These scales reflect the essence of the two mechanisms: adhesion dominates at the molecular level when the rubber enters the glassy regime, while hysteresis peaks in the rubbery–glassy transition region, where internal viscoelastic losses are maximized.
Grosch’s work thus provided the first clear experimental separation of adhesive and hysteretic contributions, and his master curves remain a cornerstone for later theories describing friction on multiscale rough surfaces, including those of Klüppel–Heinrich [114,116] and Persson [142,145].

3.2. Klüppel and Heinrich’s Friction Model

The model developed by Klüppel and Heinrich is based on the fractal scaling of rough surfaces and the linear viscoelastic response of rubber [115,116,125]. It considers both hysteresis and adhesion contributions during sliding, with explicit treatment of micro- and macro-scale roughness.
The hysteresis component is obtained by evaluating the energy dissipated through viscoelastic deformation over a spectrum of excitation frequencies activated during sliding [116]:
μ h y s ( v ) = F h y s F N = δ 2 σ 0 v ω m i n ω m a x ω E ( ω ) S ( ω ) d ω ,
where E is the loss modulus, S ( ω ) is the power spectral density of the surface, σ 0 is the nominal stress, and δ is the mean thickness of the deformed layer. The latter is proportional to the penetration depth z p , itself related to nominal stress by Hertzian arguments [62,125,177]:
δ = b z p , z p = π ξ E ( ω m i n ) σ 0 ,
where b is a scaling parameter depending on strain level (Figure 11) [178].
Thus, b is obtained from the slope of the linear fit between δ and z p at different strain levels, as shown in Figure 11.
For surfaces with two fractal scaling regimes, the integral is split into macro and micro ranges [115,125]:
μ h y s ( v ) = δ 2 σ 0 v ω m i n ω x ω E ( ω ) S ( ω ) d ω + ω x ω m a x ω E ( ω ) S ( ω ) d ω .
At high velocities, the micro-range contribution vanishes as λ m i n grows, and only macro-roughness remains effective.
The adhesive part is modeled as [114,115]:
μ a d h ( v ) = F a d h F N = τ s ( v ) σ 0 A c ( v , T , σ 0 ) A 0 ,
where A c / A 0 is the real-to-nominal contact area ratio (see Section 2.4) and τ s ( v ) is the velocity-dependent interfacial shear stress:
τ s ( v ) = τ s 0 1 + E / E 0 1 + v c v n ,
where τ s 0 = γ r s / l s , E / E 0 is the glassy-to-rubbery modulus ratio, n is linked to the relaxation spectrum exponent m by n = ( 1 m ) / ( 2 m ) [179], and v c is a critical velocity depending on T g and filler content. Persson’s theory suggests that v c v 0 ( E / E 0 ) 3 , with v 0 = a 0 / ( 2 π τ 0 ) [115], where τ 0 is the entanglement time.
Representative results show that the model reproduces the decrease in dry friction with increasing load, independently of velocity and temperature (Figure 12) [116].
In summary, Klüppel and Heinrich’s model provides a physically grounded description of dry rubber friction by combining fractal surface statistics, viscoelastic dissipation, and interfacial adhesion. It successfully captures the load and velocity dependence of measured friction, laying the basis for comparison with alternative formulations.

3.3. Persson’s Friction Model

Persson developed a multiscale theory that links the friction coefficient to sliding velocity by integrating viscoelastic energy dissipation over the full roughness spectrum [142]. Unlike the Klüppel–Heinrich model, it uses the 2D power spectral density (PSD) of surface roughness as input, thus characterizing the interface in a fully three-dimensional manner. A key ingredient is the real contact area, introduced in Section 2.4.
In sliding, the surface displacement field u i at z = 0 is defined via elastodynamics:
u i ( q , ω ) = 1 ( 2 π ) 3 d 2 x d t u i ( x , t ) e i ( q x ω t ) ,
and the hysteresis contribution follows from the fluctuating traction produced by asperities [142]:
μ h y s = 1 2 q 0 q 1 q 3 C ( q ) P ( q ) 0 2 π cos ϕ Im E * ( q v cos ϕ ) ( 1 ν 2 ) σ 0 d ϕ d q ,
where C ( q ) is the (2D) PSD, E * is the complex modulus, σ 0 is the nominal pressure, v is the sliding velocity, and ϕ is the angle between q and the sliding direction.
Incomplete contact is captured by
P ( q ) = 1 π 0 G ( q ) e x 2 / 4 d x = erf 1 2 G ( q ) ,
with
G ( q ) = 1 8 q 0 q 1 q 3 C ( q ) 0 2 π E * ( q v cos ϕ ) ( 1 ν 2 ) σ 0 2 d ϕ d q .
Water or contaminants can be incorporated by reducing the upper cutoff q 1 (filled cavities), which lowers predicted friction [142].
A central sensitivity of the model is the choice of q 1 (or magnification ζ max ): including finer scales increases friction [180].
Persson later included flash temperature effects, which modify E * through time–temperature superposition (e.g., WLF) when v 10 2 m/s [174,181,182,183,184]. The temperature field T(x,t) in the rubber satisfies
T t D T 2 T = Q ˙ ( x , t ) ρ C V ,
leading to a scale-dependent flash temperature T q (computed iteratively). In practice, the scale-dependent flash temperature T q is obtained by updating the local viscoelastic modulus E * ( ω , T q ) through time–temperature superposition until the temperature increment between iterations falls below a small threshold (typically 1–2%), without significant computational overhead for single-point evaluations. The resulting non-isothermal generalizations of (34) and (36) are
μ h y s = 1 2 q 0 q 1 q 3 C ( q ) P ( q ) 0 2 π cos ϕ Im E * ( q v cos ϕ , T q ) ( 1 ν 2 ) σ 0 d ϕ d q ,
G ( q ) = 1 8 q 0 q 1 q 3 C ( q ) 0 2 π E * ( q v cos ϕ , T q ) ( 1 ν 2 ) σ 0 2 d ϕ d q .
Physically, at intermediate magnifications (roughly ζ 2 5 ), the interface consists of spaced macro-asperity contacts that heat rapidly; at higher magnifications, dense micro-asperities within each macro-contact area interact thermally and elastically, requiring the multi-scale averaging used for T q [181].
Persson initially neglected adhesion for typical lower cutoffs (≳0.1 μ m) [142]. However, subsequent studies have attempted to investigate the adhesive contribution more thoroughly, although this aspect of friction remains largely empirical and not well understood [185], and he later proposed an empirical adhesive term similar to Klüppel’s [186]:
μ a d h = τ f p 0 A c A 0 , τ f = τ f 0 exp c log 10 v v c 2 ,
where τ f 0 is typically 5–11 MPa, c 0.1 , and v c is thermally activated:
v c = v 0 exp ϵ k B 1 T 1 T g ,
where v 0 = v 0 * exp [ ( ϵ / k B ) ( T 0 1 T g 0 1 ) ] 5 × 10 7 m/s (definitions as in the original formulation).
Several studies have revisited or leveraged Persson’s framework. Ciavarella proposed simplified formulations starting from the cutoff selection problem [146,187].
Fina et al. compared Grosch’s data [170] with Persson’s predictions: after tuning material parameters (but not q 1 ), the hysteresis peak was captured while low-/high-speed discrepancies persisted, likely due to missing adhesion and spectral differences [188].
More recently, Furlan and Mavros used an ANN to replicate Persson’s flash temperature friction model, enabling real-time friction estimation from compound operating conditions and surface spectra. Integrated into a tire and vehicle model, it captured surface-dependent variations with notable effects on braking and handling. [189].
Overall, Persson’s theory provides a rigorous multiscale connection between PSD, viscoelasticity, and operating conditions, and—since being extended to flash temperature—it has become a cornerstone for both fundamental studies and applied tire models. Its main limitations are the sensitivity to the upper cutoff (magnification) [180] and the computational cost of the non-isothermal implementation, which motivate ongoing simplifications and hybrid approaches.

4. Discussion and Research Gaps

The preceding sections reviewed the main theories describing rubber contact with rough surfaces and the resulting friction. Despite differences in complexity and assumptions, all models address how surface roughness, viscoelastic behavior, and operating conditions influence real contact area and energy dissipation. This concluding section compares the models, outlines their strengths and limitations, and identifies open research gaps.

4.1. Critical Analysis of Contact Mechanics Models

Classical approaches to contact mechanics, such as Hertzian theory, provide the analytical foundation for contact analysis but are inadequate when applied to elastomers in contact with rough surfaces. Hertz assumes smooth, perfectly elastic bodies and therefore neglects the multiscale roughness, viscoelasticity, and adhesion typical of rubber interfaces. Despite these limitations, it remains a useful baseline for comparison.
The Greenwood–Williamson (GW) model improved upon Hertz’s model by introducing a statistical description of surface roughness through an ensemble of asperities. This allowed for partial contact and roughness distributions to be considered, but simplifications such as identical spherical asperities, uncorrelated heights, and purely elastic interactions limit its applicability to rubber, especially as it omits frequency-dependent viscoelastic effects and scale transitions.
Adhesive contact theories, particularly the Johnson–Kendall–Roberts (JKR) formulation, extended Hertz by including adhesion at the contact edge. JKR is relevant for compliant systems and small scales but still assumes smooth geometries and does not incorporate roughness or viscoelastic effects. Nevertheless, it established the basis for more advanced adhesive contact models.
Modern multiscale approaches addressed these shortcomings by combining fractal surface descriptions with viscoelastic rheology. Klüppel and Heinrich introduced a framework in which an energy-based condition defines the minimum contact length scale, below which contact becomes negligible. This concept links roughness geometry and frequency-dependent rubber response, offering a more realistic picture of load support and dissipation across scales.
Persson’s theory further advanced this view by analyzing contact as a function of magnification, using the surface power spectral density (PSD) instead of discrete asperities. Large-scale features dominate at low magnification, while finer scales emerge as magnification increases. The introduction of the cutoff wavevector q 1 defines the finest roughness scale included, but the absence of a universally accepted criterion for its selection makes predictions highly sensitive to this choice.
Overall, while modern multiscale models offer more realistic descriptions of rough contact, they face challenges related to parameter identification, sensitivity to PSD cutoffs, and computational demands. Both Klüppel–Heinrich and Persson provide strategies for scale separation and contact area estimation, but with different assumptions and mathematical treatments. Comparative analysis of their predictions remains an open research topic.
Table 1 summarizes the main contact mechanics models, highlighting their respective strengths and limitations.

4.2. Comparative Discussion of Friction Models

Friction modeling has evolved from empirical observations to increasingly sophisticated multiscale theories. Grosch’s pioneering experiments [170] established the correlation between the coefficient of friction and the viscoelastic response of rubber, showing how sliding velocity and temperature govern energy dissipation. These insights laid the foundation for hysteresis-based models, highlighting the frequency dependence of friction.
Building on this, Klüppel and Heinrich [115,116] developed a semi-analytical framework that decomposes roughness into micro- and macro-asperity contributions. Their model extends Greenwood–Williamson contact mechanics to rubber sliding, integrates the power-law behavior of viscoelastic dissipation, and introduces a dynamic interaction function linking excitation frequencies to friction. While effective in reproducing the bell-shaped μ v curve, the model requires calibration through fitting parameters and treats adhesion in an empirical way.
Persson’s model [142,181] provides a more unified formulation by explicitly coupling his multiscale contact theory with energy dissipation. By incorporating the 2D power spectral density (PSD) of surfaces, it treats roughness in three dimensions and naturally accounts for scale-dependent viscoelastic losses. The later extension to flash temperature explains the friction decay at high velocities, although at the cost of significant computational effort. Adhesion was subsequently introduced in Persson’s framework, but still through empirical assumptions that limit predictive robustness.
To address computational challenges, Furlan and Mavros [189] trained artificial neural networks (ANNs) to emulate Persson’s flash temperature model. Their implementation demonstrated real-time capability, making integration into tire and vehicle simulations feasible. However, the ANN approach sacrifices interpretability, relying instead on data-driven replication of the original physics-based model.
A comparative summary of the main friction models is presented in Table 2.
Beyond conceptual differences, quantitative comparison among the reviewed models can be achieved by expressing results in terms of directly measurable quantities. In practice, the key quantities—surface power spectral density, viscoelastic moduli and cutoff wavevectors—can be obtained from surface texture (via profilometric measurements) analysis and dynamic mechanical analysis, enabling reproducible benchmarking across compounds and roughness scales. Recent progress in surface metrology, high-speed thermography, and finite-element or reduced-order simulations provides the computational tools necessary to implement these models in predictive workflows, thus facilitating their adoption by engineers and practitioners.
Under equivalent operating conditions, both frameworks predict similar qualitative trends but differ quantitatively in their sensitivity to input parameters. The Klüppel–Heinrich model, being based on discrete asperity statistics, shows a clear dependence of the A c / A 0 on the applied pressure and a monotonic reduction with increasing sliding velocity due to the glass transition of the compound. Persson’s theory, in contrast, exhibits a stronger dependence on the upper cutoff q 1 and captures nonlinear reductions of A c / A 0 with magnification, leading to slightly lower friction levels at high velocities when flash temperature effects are included.

4.3. Research Gaps and Future Challenges

Despite significant progress, several open challenges remain in the contact and friction modeling of rubber. A central issue is the definition of cutoff criteria: most advanced models depend on a cutoff wavevector or length scale, yet its selection is often empirical and strongly affects predictions. Related to this, thermo-mechanical coupling across scales and its impact on contact area and friction remain only partially understood.
Adhesion is another unresolved aspect. While minor at high speeds and loads, it dominates at low loads or in clean, dry conditions. Current models (e.g., Klüppel and Persson) treat adhesion empirically, and a rigorous, experimentally validated adhesive formulation is still lacking. Similarly, the representation of surface roughness through power spectral density (PSD) is limited by measurement resolution and data truncation, raising the need for complementary descriptors to fully characterize road textures.
Computational tractability is also a key limitation. Persson’s flash temperature model, for example, is too demanding for large-scale or real-time applications, while surrogate models improve efficiency at the cost of physical interpretability. Furthermore, most models are calibrated on restricted conditions (e.g., sandpaper, single compounds, narrow velocity ranges), limiting generalization to realistic surfaces, multiple compounds, or variable environments. The absence of extensive experimental benchmarks hampers validation.
Future developments should aim to achieve a more quantitative and physically grounded understanding of adhesion under realistic tire–road conditions. This will likely require in situ diagnostics capable of capturing interfacial temperature, stress, and detachment dynamics, together with multiscale models that incorporate such data in a physically consistent way. Coupling experimental evidence with data-driven techniques may ultimately provide a pathway toward predictive adhesive friction. Nevertheless, it should be emphasized that the quantification of the adhesive contribution remains a topic of active debate within the tribology community and is still far less understood than the hysteretic counterpart.
In summary, current theories—from classical contact to multiscale viscoelastic frameworks—have greatly advanced the understanding of tire–road interaction. Yet, major research gaps persist regarding multiscale coupling, adhesion, roughness representation, and computational efficiency. Future progress will depend on integrating high-resolution experimental methods, improved surface characterization, hybrid physics–data-driven approaches, and validation within full vehicle simulations, thus bridging the gap between academic insight and industrial applicability.

Author Contributions

Conceptualization, R.S.; Methodology, R.S. and G.F.; Formal Analysis, R.S. and A.G.; Investigation, R.S., G.N.D. and A.S.; Resources, R.S., G.N.D. and A.S.; Data Curation, R.S. and G.F.; Writing—original draft preparation, R.S.; writing—review and editing, G.F. and F.F.; Visualization, R.S., G.F. and A.G.; Supervision, A.S., F.T. and F.F.; Project Administration, F.T. and F.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partly supported by the project “FINGERTIPS—Friction-Increase Nanopoles and Grooves Exploitation in Road-Tire Interaction for Preventive Safety’’ funded by the Italian MUR ‘‘Progetti di Ricerca di Rilevante Interesse Nazionale (PRIN 2022 PNRR PE8)’’—grant n. P2022LZWF5—P.I. Flavio Farroni.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Johnson, K.L.; Johnson, K.L. Contact Mechanics; Cambridge University Press: Cambridge, UK, 1987. [Google Scholar]
  2. Alvarez, L.; Yi, J.; Horowitz, R.; Olmos, L. Dynamic friction model-based tire-road friction estimation and emergency braking control. J. Dyn. Sys. Meas. Control 2005, 127, 22–32. [Google Scholar] [CrossRef]
  3. Barber, J.R.; Ciavarella, M. Contact mechanics. Int. J. Solids Struct. 2000, 37, 29–43. [Google Scholar] [CrossRef]
  4. van Kuilenburg, J.; Masen, M.A.; van der Heide, E. A review of fingerpad contact mechanics and friction and how this affects tactile perception. Proc. Inst. Mech. Eng. Part J J. Eng. Tribol. 2015, 229, 243–258. [Google Scholar] [CrossRef]
  5. Barbaro, M.; Genovese, A.; Timpone, F.; Sakhnevych, A. Extension of the multiphysical magic formula tire model for ride comfort applications. Nonlinear Dyn. 2024, 112, 4183–4208. [Google Scholar] [CrossRef]
  6. Dutkiewicz, M.; Velychkovych, A.; Shatskyi, I.; Shopa, V. Efficient model of the interaction of elastomeric filler with an open shell and a chrome-plated shaft in a dry friction damper. Materials 2022, 15, 4671. [Google Scholar] [CrossRef]
  7. Shatskyi, I.; Velychkovych, A. Analytical model of structural damping in friction module of shell shock absorber connected to spring. Shock Vib. 2023, 2023, 4140583. [Google Scholar] [CrossRef]
  8. Velychkovych, A.; Mykhailiuk, V.; Andrusyak, A. Numerical model for studying the properties of a new friction damper developed based on the shell with a helical cut. Appl. Mech. 2025, 6, 1. [Google Scholar] [CrossRef]
  9. Velychkovych, A.; Mykhailiuk, V.; Andrusyak, A. Evaluation of the Adaptive Behavior of a Shell-Type Elastic Element of a Drilling Shock Absorber with Increasing External Load Amplitude. Vibration 2025, 8, 60. [Google Scholar] [CrossRef]
  10. Shull, K.R. Contact mechanics and the adhesion of soft solids. Mater. Sci. Eng. R Rep. 2002, 36, 1–45. [Google Scholar] [CrossRef]
  11. Bhushan, B. Contact mechanics of rough surfaces in tribology: Single asperity contact. Appl. Mech. Rev. 1996, 49, 275–298. [Google Scholar] [CrossRef]
  12. Jean, M. The non-smooth contact dynamics method. Comput. Methods Appl. Mech. Eng. 1999, 177, 235–257. [Google Scholar] [CrossRef]
  13. Tichy, J.A.; Meyer, D.M. Review of solid mechanics in tribology. Int. J. Solids Struct. 2000, 37, 391–400. [Google Scholar] [CrossRef]
  14. Persson, B. Rubber friction and tire dynamics. J. Phys. Condens. Matter 2010, 23, 015003. [Google Scholar] [CrossRef]
  15. Gravante, G.; Dell’Annunziata, G.N.; Barbaro, M.; Farroni, F. Experimental Methodology for Tire Ride Analysis Based on Outdoor Cleat Testing. SAE Int. J. Veh. Dyn. Stab. NVH 2025, 9, 357–376. [Google Scholar] [CrossRef]
  16. Ghaednia, H.; Wang, X.; Saha, S.; Xu, Y.; Sharma, A.; Jackson, R.L. A review of elastic–plastic contact mechanics. Appl. Mech. Rev. 2017, 69, 060804. [Google Scholar] [CrossRef]
  17. Jackson, R.L.; Ghaednia, H.; Lee, H.; Rostami, A.; Wang, X. Contact mechanics. In Tribology for Scientists and Engineers: From Basics to Advanced Concepts; Springer: New York, NY, USA, 2013; pp. 93–140. [Google Scholar]
  18. Goryacheva, I.G. Contact Mechanics in Tribology; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2013; Volume 61. [Google Scholar]
  19. Zhang, S.; Li, D.; Liu, Y. Friction behavior of rough surfaces on the basis of contact mechanics: A review and prospects. Micromachines 2022, 13, 1907. [Google Scholar] [CrossRef] [PubMed]
  20. Popov, V.L. Contact Mechanics and Friction; Springer: Berlin/Heidelberg, Germany, 2010. [Google Scholar]
  21. Popov, V.L.; Heß, M.; Willert, E. Handbook of Contact Mechanics: Exact Solutions of Axisymmetric Contact Problems; Springer Nature: Berlin/Heidelberg, Germany, 2019. [Google Scholar]
  22. Raous, M.; Jean, M.; Moreau, J.J. Contact Mechanics; Plenum Press: New York, NY, USA, 1995. [Google Scholar]
  23. Gilardi, G.; Sharf, I. Literature survey of contact dynamics modelling. Mech. Mach. Theory 2002, 37, 1213–1239. [Google Scholar] [CrossRef]
  24. Eid, H.; Pamp, A. An Introductory Overview of Contact Mechanics and Adhesion; Springer: Berlin/Heidelberg, Germany, 2006. [Google Scholar]
  25. Ciavarella, M.; Joe, J.; Papangelo, A.; Barber, J. The role of adhesion in contact mechanics. J. R. Soc. Interface 2019, 16, 20180738. [Google Scholar] [CrossRef]
  26. Fischer-Cripps, A.C. Introduction to Contact Mechanics; Springer: New York, NY, USA, 2000. [Google Scholar]
  27. Johnson, K. Mechanics of adhesion. Tribol. Int. 1998, 31, 413–418. [Google Scholar] [CrossRef]
  28. Roberts, A.D. Rubber contact phenomena. Rubber Chem. Technol. 2014, 87, 383–416. [Google Scholar] [CrossRef]
  29. Dubois, G.; Cesbron, J.; Yin, H.P.; Anfosso-Lédée, F. Numerical evaluation of tyre/road contact pressures using a multi-asperity approach. Int. J. Mech. Sci. 2012, 54, 84–94. [Google Scholar] [CrossRef]
  30. Schallamach, A. Friction and abrasion of rubber. Wear 1958, 1, 384–417. [Google Scholar] [CrossRef]
  31. Müser, M.H.; Dapp, W.B.; Bugnicourt, R.; Sainsot, P.; Lesaffre, N.; Lubrecht, T.A.; Persson, B.N.; Harris, K.; Bennett, A.; Schulze, K.; et al. Meeting the contact-mechanics challenge. Tribol. Lett. 2017, 65, 118. [Google Scholar] [CrossRef]
  32. Mao, K.; Sun, Y.; Bell, T. Contact mechanics of engineering surfaces: State of the art. Surf. Eng. 1994, 10, 297–306. [Google Scholar] [CrossRef]
  33. Hertz, H. Über die Berührung fester elastischer Körper. J. Reine Angew. Math. 1881, 92, 156. [Google Scholar]
  34. Fischer-Cripps, A. The Hertzian contact surface. J. Mater. Sci. 1999, 34, 129–137. [Google Scholar] [CrossRef]
  35. Johnson, K.L. One hundred years of Hertz contact. Proc. Inst. Mech. Eng. 1982, 196, 363–378. [Google Scholar] [CrossRef]
  36. Wang, Q.J.; Zhu, D. Hertz theory: Contact of spherical surfaces. In Encyclopedia of Tribology; Springer: Boston, MA, USA, 2013; pp. 1654–1662. [Google Scholar]
  37. Tripp, J.H. Hertzian Contact in Two and Three Dimensions; National Aeronautics and Space Administration, Scientific and Technical Information Branch: Washington, DC, USA, 1985. [Google Scholar]
  38. Etter, J. Tutorial Report on Hertzian Contact Stress Theory; University of Arizona: Tucson, AZ, USA, 2000. [Google Scholar]
  39. Purushothaman, P.; Thankachan, P.; John, S.; Purushothaman, M. Hertz contact stress analysis and validation using finite element analysis. Int. J. Res. Appl. Sci. Eng. Technol. 2014, 2, 531–538. [Google Scholar]
  40. Dintwa, E.; Tijskens, E.; Ramon, H. On the accuracy of the Hertz model to describe the normal contact of soft elastic spheres. Granul. Matter 2008, 10, 209–221. [Google Scholar] [CrossRef]
  41. Sneddon, I.N. The relation between load and penetration in the axisymmetric Boussinesq problem for a punch of arbitrary profile. Int. J. Eng. Sci. 1965, 3, 47–57. [Google Scholar] [CrossRef]
  42. Galin, L. Kontakn’e Zadachi Teorii Uprugosti; Gosudarstvennoe Izdatel’stvo Tekhniko-Teoreticheskoi Literatury: Moscow, Russia, 1953. [Google Scholar]
  43. Bhushan, B. Contact mechanics of rough surfaces in tribology: Multiple asperity contact. Tribol. Lett. 1998, 4, 1–35. [Google Scholar] [CrossRef]
  44. Taylor, R.I. Rough surface contact modelling—A review. Lubricants 2022, 10, 98. [Google Scholar] [CrossRef]
  45. Majumdar, A.; Bhushan, B. Characterization and modeling of surface roughness and contact mechanics. In Handbook of Micro/Nano Tribology; CRC Press: Boca Raton, FL, USA, 1995; pp. 109–165. [Google Scholar]
  46. Sofonea, M.; Matei, A. Mathematical Models in Contact Mechanics; Cambridge University Press: Cambridge, UK, 2012. [Google Scholar]
  47. Bowden, F.P.; Tabor, D. The area of contact between stationary and moving surfaces. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1939, 169, 391–413. [Google Scholar]
  48. Bowden, F.; Tabor, D. Friction, lubrication and wear: A survey of work during the last decade. Br. J. Appl. Phys. 1966, 17, 1521. [Google Scholar] [CrossRef]
  49. Bowden, F.P.; Tabor, D. The Friction and Lubrication of Solids; Oxford University Press: Oxford, UK, 2001; Volume 1. [Google Scholar]
  50. Tabor, D. Friction, surface science and tribology—A personal view. Proc. Inst. Mech. Eng. Part C Mech. Eng. Sci. 1991, 205, 365–378. [Google Scholar] [CrossRef]
  51. Carbone, G.; Bottiglione, F. Contact mechanics of rough surfaces: A comparison between theories. Meccanica 2011, 46, 557–565. [Google Scholar] [CrossRef]
  52. Archard, J. Contact and rubbing of flat surfaces. J. Appl. Phys. 1953, 24, 981–988. [Google Scholar] [CrossRef]
  53. Archard, J. Elastic deformation and the laws of friction. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1957, 243, 190–205. [Google Scholar]
  54. Archard, J. Single contacts and multiple encounters. J. Appl. Phys. 1961, 32, 1420–1425. [Google Scholar] [CrossRef]
  55. Emami, A.; Khaleghian, S.; Taheri, S. Asperity-based modification on theory of contact mechanics and rubber friction for self-affine fractal surfaces. Friction 2021, 9, 1707–1725. [Google Scholar] [CrossRef]
  56. Adams, G.G.; Mu ftu, S.; Azhar, N.M. A scale-dependent model for multi-asperity contact and friction. J. Tribol. 2003, 125, 700–708. [Google Scholar] [CrossRef]
  57. Popov, V.L.; Dimaki, A.; Psakhie, S.; Popov, M. On the role of scales in contact mechanics and friction between elastomers and randomly rough self-affine surfaces. Sci. Rep. 2015, 5, 11139. [Google Scholar] [CrossRef]
  58. Hanaor, D.A.; Gan, Y.; Einav, I. Contact mechanics of fractal surfaces by spline assisted discretisation. Int. J. Solids Struct. 2015, 59, 121–131. [Google Scholar] [CrossRef]
  59. Buzio, R.; Boragno, C.; Biscarini, F.; Buatier De Mongeot, F.; Valbusa, U. The contact mechanics of fractal surfaces. Nat. Mater. 2003, 2, 233–236. [Google Scholar] [CrossRef]
  60. Borodich, F.M. Fractal contact mechanics. In Encyclopedia of Tribology; Springer: Boston, MA, USA, 2013; pp. 1249–1258. [Google Scholar]
  61. Majumdar, A.; Bhushan, B. Role of fractal geometry in roughness characterization and contact mechanics of surfaces. J. Tribol. 1990, 112, 205–216. [Google Scholar] [CrossRef]
  62. Greenwood, J.A.; Williamson, J.P. Contact of nominally flat surfaces. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1966, 295, 300–319. [Google Scholar]
  63. Bush, A.; Gibson, R.; Thomas, T. The elastic contact of a rough surface. Wear 1975, 35, 87–111. [Google Scholar] [CrossRef]
  64. Ciavarella, M.; Greenwood, J.; Paggi, M. Inclusion of “interaction” in the Greenwood and Williamson contact theory. Wear 2008, 265, 729–734. [Google Scholar] [CrossRef]
  65. Ciavarella, M.; Delfine, V.; Demelio, G. A “re-vitalized” Greenwood and Williamson model of elastic contact between fractal surfaces. J. Mech. Phys. Solids 2006, 54, 2569–2591. [Google Scholar] [CrossRef]
  66. Adams, G.; Nosonovsky, M. Contact modeling—Forces. Tribol. Int. 2000, 33, 431–442. [Google Scholar] [CrossRef]
  67. Yakovenko, A.A.; Goryacheva, I.G. Analysis of the discrete contact characteristics based on the Greenwood-Williamson model and the localization principle. Friction 2024, 12, 1042–1056. [Google Scholar] [CrossRef]
  68. Ciavarella, M.; Leoci, F. An assessment of the Greenwood-Williamson and other asperities models, with special reference to electrical conductance. J. Tribol. 2006, 128, 10–17. [Google Scholar] [CrossRef]
  69. Farroni, F.; Stefanelli, R.; Dell’Annunziata, G.N.; Timpone, F. Impact of the tire/road local contact area on tire friction, with a novel approach based on Greenwood-Williamson theory and validation in both laboratory and outdoor racing environments. Tribol. Int. 2025, 210, 110772. [Google Scholar] [CrossRef]
  70. Zhao, T.; Feng, Y. Extended Greenwood–Williamson models for rough spheres. J. Appl. Mech. 2018, 85, 101007. [Google Scholar] [CrossRef]
  71. Li, C.Y.; Ding, Y.; Liang, X.M.; Wang, G.F. An improved elastic–plastic contact model with asperity interactions based on Greenwood–Williamson theory. Acta Mech. 2023, 234, 5187–5201. [Google Scholar] [CrossRef]
  72. McWaid, T.; Marschall, E. Application of the modified Greenwood and Williamson contact model for the prediction of thermal contact resistance. Wear 1992, 152, 263–277. [Google Scholar] [CrossRef]
  73. Jones, R.E. A Greenwood-Williamson model of small-scale friction. J. Appl. Mech. 2007, 74, 31–40. [Google Scholar] [CrossRef]
  74. Bush, A.; Gibson, R.; Keogh, G. The limit of elastic deformation in the contact of rough surfaces. Mech. Res. Commun. 1976, 3, 169–174. [Google Scholar] [CrossRef]
  75. Zhao, Y.P.; Wang, L.; Yu, T. Mechanics of adhesion in MEMS—A review. J. Adhes. Sci. Technol. 2003, 17, 519–546. [Google Scholar] [CrossRef]
  76. Persson, B.N.; Albohr, O.; Tartaglino, U.; Volokitin, A.I.; Tosatti, E. On the nature of surface roughness with application to contact mechanics, sealing, rubberfriction and adhesion. J. Phys. Condens. Matter 2004, 17, R1. [Google Scholar] [CrossRef]
  77. Popov, V.L.; Li, Q.; Lyashenko, I.A. Contact mechanics and friction: Role of adhesion. Friction 2025, 13, 9440964. [Google Scholar] [CrossRef]
  78. Marshall, S.J.; Bayne, S.C.; Baier, R.; Tomsia, A.P.; Marshall, G.W. A review of adhesion science. Dent. Mater. 2010, 26, e11–e16. [Google Scholar] [CrossRef]
  79. Adamson, A.W.; Gast, A.P. Physical Chemistry of Surfaces; Interscience Publishers: New York, NY, USA, 1967; Volume 150. [Google Scholar]
  80. Lee, L.H. Fundamentals of Adhesion; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  81. Bhushan, B. Adhesion and stiction: Mechanisms, measurement techniques, and methods for reduction. J. Vac. Sci. Technol. B Microelectron. Nanometer Struct. Processing, Meas. Phenom. 2003, 21, 2262–2296. [Google Scholar] [CrossRef]
  82. Kendall, K.; Roberts, A. van der Waals forces influencing adhesion of cells. Philos. Trans. R. Soc. B Biol. Sci. 2015, 370, 20140078. [Google Scholar] [CrossRef]
  83. Roberts, A. A guide to estimating the friction of rubber. Rubber Chem. Technol. 1992, 65, 673–686. [Google Scholar] [CrossRef]
  84. Kendall, K. Adhesion: Molecules and mechanics. Science 1994, 263, 1720–1725. [Google Scholar] [CrossRef]
  85. Bradley, R.S. LXXIX. The cohesive force between solid surfaces and the surface energy of solids. Lond. Edinb. Dublin Philos. Mag. J. Sci. 1932, 13, 853–862. [Google Scholar] [CrossRef]
  86. Nakajima, Y. Advanced Tire Mechanics; Springer: Berlin/Heidelberg, Germany, 2019. [Google Scholar]
  87. Schultz, J.; Nardin, M. Theories and mechanisms of adhesion. Mater. Eng. N. Y. 1999, 14, 1–26. [Google Scholar]
  88. Packham, D. The mechanical theory of adhesion. In Handbook of Adhesive Technology; CRC Press: Boca Raton, FL, USA, 2003; pp. 69–93. [Google Scholar]
  89. Johnson, K.L.; Kendall, K.; Roberts, A. Surface energy and the contact of elastic solids. Proc. R. Soc. Lond. A Math. Phys. Sci. 1971, 324, 301–313. [Google Scholar]
  90. Derjaguin, B.V.; Muller, V.M.; Toporov, Y.P. Effect of contact deformations on the adhesion of particles. J. Colloid Interface Sci. 1975, 53, 314–326. [Google Scholar] [CrossRef]
  91. Barthel, E. Adhesive elastic contacts: JKR and more. J. Phys. D Appl. Phys. 2008, 41, 163001. [Google Scholar] [CrossRef]
  92. Hui, C.Y.; Baney, J.; Kramer, E. Contact mechanics and adhesion of viscoelastic spheres. Langmuir 1998, 14, 6570–6578. [Google Scholar] [CrossRef]
  93. Coetzee, C. Simplified johnson-kendall-roberts (sjkr) contact model. In University of Stellenbosch Report; University of Stellenbosch: Stellenbosch, South Africa, 2020. [Google Scholar]
  94. Charrault, E.; Gauthier, C.; Marie, P.; Schirrer, R. Experimental and theoretical analysis of a dynamic JKR contact. Langmuir 2009, 25, 5847–5854. [Google Scholar] [CrossRef]
  95. Johnson, K. The adhesion of two elastic bodies with slightly wavy surfaces. Int. J. Solids Struct. 1995, 32, 423–430. [Google Scholar] [CrossRef]
  96. Pashley, M. Further consideration of the DMT model for elastic contact. Colloids Surf. 1984, 12, 69–77. [Google Scholar] [CrossRef]
  97. Violano, G.; Demelio, G.P.; Afferrante, L. On the DMT adhesion theory: From the first studies to the modern applications in rough contacts. Procedia Struct. Integr. 2018, 12, 58–70. [Google Scholar] [CrossRef]
  98. Violano, G.; Afferrante, L. On DMT methods to calculate adhesion in rough contacts. Tribol. Int. 2019, 130, 36–42. [Google Scholar] [CrossRef]
  99. Greenwood, J. On the DMT theory. Tribol. Lett. 2007, 26, 203–211. [Google Scholar] [CrossRef]
  100. Prokopovich, P.; Perni, S. Comparison of JKR-and DMT-based multi-asperity adhesion model: Theory and experiment. Colloids Surf. A Physicochem. Eng. Asp. 2011, 383, 95–101. [Google Scholar] [CrossRef]
  101. Tabor, D. Surface forces and surface interactions. In Plenary and Invited Lectures; Elsevier: Amsterdam, The Netherlands, 1977; pp. 3–14. [Google Scholar]
  102. Muller, V.; Yushchenko, V.; Derjaguin, B. On the influence of molecular forces on the deformation of an elastic sphere and its sticking to a rigid plane. J. Colloid Interface Sci. 1980, 77, 91–101. [Google Scholar] [CrossRef]
  103. Maugis, D. Adhesion of spheres: The JKR-DMT transition using a Dugdale model. J. Colloid Interface Sci. 1992, 150, 243–269. [Google Scholar] [CrossRef]
  104. Zhang, Y. Transitions between different contact models. J. Adhes. Sci. Technol. 2008, 22, 699–715. [Google Scholar]
  105. Maugis, D.; Gauthier-Manuel, B. JKR-DMT transition in the presence of a liquid meniscus. J. Adhes. Sci. Technol. 1994, 8, 1311–1322. [Google Scholar] [CrossRef]
  106. Maugis, D. The JKR-DMT transition in the presence of a liquid meniscus and the extension of the JKR theory to large contact radii. In Contact Mechanics; Springer: Boston, MA, USA, 1995; pp. 335–342. [Google Scholar]
  107. Maugis, D. On the contact and adhesion of rough surfaces. J. Adhes. Sci. Technol. 1996, 10, 161–175. [Google Scholar] [CrossRef]
  108. Maugis, D. Contact, Adhesion and Rupture of Elastic Solids; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2000; Volume 130. [Google Scholar]
  109. Shi, X.; Zhao, Y.P. Comparison of various adhesion contact theories and the influence of dimensionless load parameter. J. Adhes. Sci. Technol. 2004, 18, 55–68. [Google Scholar] [CrossRef]
  110. Ramakrishna, S.N.; Nalam, P.C.; Clasohm, L.Y.; Spencer, N.D. Study of adhesion and friction properties on a nanoparticle gradient surface: Transition from JKR to DMT contact mechanics. Langmuir 2013, 29, 175–182. [Google Scholar] [CrossRef]
  111. Carbone, G.; Lorenz, B.; Persson, B.N.; Wohlers, A. Contact mechanics and rubber friction for randomly rough surfaces with anisotropic statistical properties. Eur. Phys. J. E 2009, 29, 275–284. [Google Scholar] [CrossRef]
  112. Scaraggi, M.; Persson, B.N. General contact mechanics theory for randomly rough surfaces with application to rubber friction. J. Chem. Phys. 2015, 143, 224111. [Google Scholar] [CrossRef]
  113. Flores, P. Contact mechanics for dynamical systems: A comprehensive review. Multibody Syst. Dyn. 2022, 54, 127–177. [Google Scholar] [CrossRef]
  114. Le Gal, A.; Yang, X.; Klüppel, M. Evaluation of sliding friction and contact mechanics of elastomers based on dynamic-mechanical analysis. J. Chem. Phys. 2005, 123, 014704. [Google Scholar] [CrossRef]
  115. Kluppel, M.; Heinrich, G. Rubber friction on self-affine road tracks. Rubber Chem. Technol. 2000, 73, 578–606. [Google Scholar] [CrossRef]
  116. Lang, A.; Klüppel, M. Influences of temperature and load on the dry friction behaviour of tire tread compounds in contact with rough granite. Wear 2017, 380, 15–25. [Google Scholar] [CrossRef]
  117. Heinrich, G.; Klüppel, M. Rubber friction, tread deformation and tire traction. Wear 2008, 265, 1052–1060. [Google Scholar] [CrossRef]
  118. Le Gal, A.; Klüppel, M. Modeling of rubber friction: A quantitative description of the hysteresis and adhesion contribution. In Constitutive Models for Rubber IV; Routledge: Oxfordshire, UK, 2017; pp. 509–514. [Google Scholar]
  119. Lang, A.; Klüppel, M. Modelling predictions and experimental investigations of rubber friction and tire traction. In Proceedings of the 4th International Tyre Colloquium: Tyre Models for Vehicle Dynamics Analysis, Guildford, UK, 20–21 April 2015; University of Surrey: Guildford, UK, 2015; pp. 228–233. [Google Scholar]
  120. Li, L.; Wang, K.C.; Luo, W. Pavement friction estimation based on the Heinrich/Klüppel model. Period. Polytech. Transp. Eng. 2016, 44, 89–96. [Google Scholar] [CrossRef]
  121. Lang, A.; Klüppel, M. Temperature and Pressure dependence of the friction properties of tire tread compounds on rough granite. In Proceedings of the KHK 11th Fall Rubber Colloquium, Hannover, Germany, 26–28 November 2014. [Google Scholar]
  122. Klüppel, M.; Müller, A.; Le Gal, A.; Heinrich, G.; Continental, A. Dynamic Contact of Tires Dynamic Contact of Tires with Road Tracks with Road Tracks; Rubber Division, American Chemical Society: San Francisco, CA, USA, 2003. [Google Scholar]
  123. Le Gal, A.; Yang, X.; Klüppel, M. Sliding friction and contact mechanics of elastomers on rough surfaces. In Analysis and Simulation of Contact Problems; Springer: Berlin/Heidelberg, Germany, 2006; pp. 253–260. [Google Scholar]
  124. Le Gal, A.; Guy, L.; Orange, G.; Bomal, Y.; Klüppel, M. Modelling of sliding friction for carbon black and silica filled elastomers on road tracks. Wear 2008, 264, 606–615. [Google Scholar] [CrossRef]
  125. Le Gal, A.; Klüppel, M. Investigation and modelling of rubber stationary friction on rough surfaces. J. Phys. Condens. Matter 2007, 20, 015007. [Google Scholar] [CrossRef]
  126. Christensen, R.M. Theory of Viscoelasticity; Courier Corporation: North Chelmsford, MA, USA, 2013. [Google Scholar]
  127. Lakes, R.S. Viscoelastic Materials; Cambridge University Press: Cambridge, UK, 2009. [Google Scholar]
  128. Papazian, H.S. The Response of Linear Viscoelastic Materials in the Frequency Domain; The Ohio State University: Columbus, OH, USA, 1961. [Google Scholar]
  129. Schapery, R.A. On the characterization of nonlinear viscoelastic materials. Polym. Eng. Sci. 1969, 9, 295–310. [Google Scholar] [CrossRef]
  130. Tomanik, E.; Chacon, H.; Teixeira, G. A simple numerical procedure to calculate the input data of Greenwood-Williamson model of asperity contact for actual engineering surfaces. In Tribology Series; Elsevier: Amsterdam, The Netherlands, 2003; Volume 41, pp. 205–215. [Google Scholar]
  131. Nayak, P.R. Random process model of rough surfaces. J. Lubr. Technol. 1971, 93, 398–407. [Google Scholar] [CrossRef]
  132. Heinrich, G.; Klüppel, M.; Vilgis, T.A. Evaluation of self-affine surfaces and their implication for frictional dynamics as illustrated with a Rouse material. Comput. Theor. Polym. Sci. 2000, 10, 53–61. [Google Scholar] [CrossRef]
  133. Vanormelingen, K.; Degroote, B.; Vantomme, A. Quantitative characterization of the surface morphology using a height difference correlation function. J. Vac. Sci. Technol. B Microelectron. Nanometer Struct. Process. Meas. Phenom. 2006, 24, 725–729. [Google Scholar] [CrossRef]
  134. Torbruegge, S.; Wies, B. Characterization of pavement texture by means of height difference correlation and relation to wet skid resistance. J. Traffic Transp. Eng. (Engl. Ed.) 2015, 2, 59–67. [Google Scholar] [CrossRef]
  135. Landini, G.; Rigaut, J.P. A method for estimating the dimension of asymptotic fractal sets. Bioimaging 1997, 5, 65–70. [Google Scholar] [CrossRef]
  136. Persson, B.N. On the theory of rubber friction. Surf. Sci. 1998, 401, 445–454. [Google Scholar] [CrossRef]
  137. Persson, B.; Volokitin, A. Rubber friction on smooth surfaces. Eur. Phys. J. E 2006, 21, 69–80. [Google Scholar] [CrossRef]
  138. Yang, C.; Persson, B.N. Contact mechanics: Contact area and interfacial separation from small contact to fullcontact. J. Phys. Condens. Matter 2008, 20, 215214. [Google Scholar] [CrossRef]
  139. Persson, B.N. Sliding Friction: Physical Principles and Applications; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  140. Persson, B.N. Contact mechanics for randomly rough surfaces. Surf. Sci. Rep. 2006, 61, 201–227. [Google Scholar] [CrossRef]
  141. Persson, B.N.; Tosatti, E. Physics of Sliding Friction; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2013; Volume 311. [Google Scholar]
  142. Persson, B.N. Theory of rubber friction and contact mechanics. J. Chem. Phys. 2001, 115, 3840–3861. [Google Scholar] [CrossRef]
  143. Persson, B. Adhesion between an elastic body and a randomly rough hard surface. Eur. Phys. J. E 2002, 8, 385–401. [Google Scholar] [CrossRef]
  144. Jacobs, T.D.; Junge, T.; Pastewka, L. Quantitative characterization of surface topography using spectral analysis. Surf. Topogr. Metrol. Prop. 2017, 5, 013001. [Google Scholar] [CrossRef]
  145. Lorenz, B.; Persson, B.; Fortunato, G.; Giustiniano, M.; Baldoni, F. Rubber friction for tire tread compound on road surfaces. J. Phys. Condens. Matter 2013, 25, 095007. [Google Scholar] [CrossRef]
  146. Genovese, A.; Farroni, F.; Papangelo, A.; Ciavarella, M. A discussion on present theories of rubber friction, with particular reference to different possible choices of arbitrary roughness cutoff parameters. Lubricants 2019, 7, 85. [Google Scholar] [CrossRef]
  147. Sanyal, A. On the Short Wavelength Cutoff Limit in Persson’s Hysteresis Friction Theory. J. Tribol. 2026, 148, 014501. [Google Scholar] [CrossRef]
  148. Müser, M.H.; Dapp, W.B. The contact mechanics challenge: Problem definition. arXiv 2015, arXiv:1512.02403. [Google Scholar] [CrossRef]
  149. Müser, M.H.; Nicola, L. Modeling the surface topography dependence of friction, adhesion, and contact compliance. MRS Bull. 2022, 47, 1221–1228. [Google Scholar] [CrossRef]
  150. Farroni, F. Development of a Grip and Thermodynamics Sensitive Tyre/Road Interaction Forces Characterization Procedure Employed in High-Performance Vehicles Simulation; University of Naples Federico” II: Naples, Italy, 2014. [Google Scholar]
  151. Pennestrì, E.; Rossi, V.; Salvini, P.; Valentini, P.P. Review and comparison of dry friction force models. Nonlinear Dyn. 2016, 83, 1785–1801. [Google Scholar] [CrossRef]
  152. Van Der Steen, R. Tyre/Road Friction Modeling: Literature Survey; Eindhoven University of Technology: Eindhoven, The Netherlands, 2007. [Google Scholar]
  153. Persson, B.; Xu, R. Rubber Friction: Theory, Mechanisms, and Challenges. arXiv 2025, arXiv:2507.18782. [Google Scholar] [CrossRef]
  154. KUMMER, H.W. Rubber and Tire Friction; The Pennsylvania State University: University Park, PA, USA, 1965. [Google Scholar]
  155. Smith, R.H. Analyzing Friction in the Design of Rubber Products and Their Paired Surfaces; CRC Press: Boca Raton, FL, USA, 2008. [Google Scholar]
  156. Greenwood, J.; Minshall, H.; Tabor, D. Hysteresis losses in rolling and sliding friction. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1961, 259, 480–507. [Google Scholar]
  157. Moore, D.; Geyer, W. A review of hysteresis theories for elastomers. Wear 1974, 30, 1–34. [Google Scholar] [CrossRef]
  158. Heinrich, G. Hysteresis friction of sliding rubbers on rough and fractal surfaces. Rubber Chem. Technol. 1997, 70, 1–14. [Google Scholar] [CrossRef]
  159. Wojewoda, J.; Stefański, A.; Wiercigroch, M.; Kapitaniak, T. Hysteretic effects of dry friction: Modelling and experimental studies. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2008, 366, 747–765. [Google Scholar] [CrossRef]
  160. McFarlane, J.; Tabor, D. Relation between friction and adhesion. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1950, 202, 244–253. [Google Scholar]
  161. Chernyak, Y.B.; Leonov, A. On the theory of the adhesive friction of elastomers. Wear 1986, 108, 105–138. [Google Scholar] [CrossRef]
  162. Zeng, H. Polymer Adhesion, Friction, and Lubrication; Wiley: Hoboken, NJ, USA, 2013. [Google Scholar]
  163. Landheer, D.; De Gee, A. Adhesion, friction, and wear. MRS Bull. 1991, 16, 36–40. [Google Scholar] [CrossRef]
  164. Chaudhury, M.K.; Owen, M.J. Adhesion hysteresis and friction. Langmuir 1993, 9, 29–31. [Google Scholar] [CrossRef]
  165. Israelachvili, J.N.; Chen, Y.L.; Yoshizawa, H. Relationship between adhesion and friction forces. J. Adhes. Sci. Technol. 1994, 8, 1231–1249. [Google Scholar] [CrossRef]
  166. Popov, V.L. Adhesive contribution to friction. In Proceedings of the International Conference on Advanced Materials with Hierarchical Structure for New Technologies and Reliable Structures 2019, Tomsk, Russia, 1–5 October 2019; AIP Publishing LLC: Melville, NY, USA, 2019; Volume 2167, p. 020286. [Google Scholar]
  167. Savkoor, A. On the friction of rubber. Wear 1965, 8, 222–237. [Google Scholar] [CrossRef]
  168. Persson, B.N.; Tosatti, E. Qualitative theory of rubber friction and wear. J. Chem. Phys. 2000, 112, 2021–2029. [Google Scholar] [CrossRef]
  169. Moore, D.F. Friction and wear in rubbers and tyres. Wear 1980, 61, 273–282. [Google Scholar] [CrossRef]
  170. Grosch, K. The relation between the friction and visco-elastic properties of rubber. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1963, 274, 21–39. [Google Scholar]
  171. Grosch, K. The rolling resistance, wear and traction properties of tread compounds. Rubber Chem. Technol. 1996, 69, 495–568. [Google Scholar] [CrossRef]
  172. Grosch, K. Rubber friction and tire traction. In The Pneumatic Tire; 2006; p. 421. Available online: https://andrzejdebowski.wat.edu.pl/include/wiedza/kolo-ogumione/The%20Pneumatic%20Tire%20NHTSA.pdf (accessed on 27 October 2025).
  173. Grosch, K. Visco-elastic properties and the friction of solids: Relation between the friction and visco-elastic properties of rubber. Nature 1963, 197, 858–859. [Google Scholar] [CrossRef]
  174. Williams, M.L.; Landel, R.F.; Ferry, J.D. The temperature dependence of relaxation mechanisms in amorphous polymers and other glass-forming liquids. J. Am. Chem. Soc. 1955, 77, 3701–3707. [Google Scholar] [CrossRef]
  175. Dagdug, L.; García-Colín, L. Generalization of the Williams–Landel–Ferry equation. Phys. A Stat. Mech. Its Appl. 1998, 250, 133–141. [Google Scholar] [CrossRef]
  176. Williams, M.L.; Ferry, J.D. Second approximation calculations of mechanical and electrical relaxation and retardation distributions. J. Polym. Sci. 1953, 11, 169–175. [Google Scholar] [CrossRef]
  177. Hertz, H. Miscellaneous Papers; Macmillan: New York, NY, USA, 1896. [Google Scholar]
  178. Le Gal, A. Investigation and modelling of adhesion friction on rough surfaces. Kautshcuk Gummi Kunststoffe 2006, 59, 308. [Google Scholar]
  179. Persson, B.; Brener, E. Crack propagation in viscoelastic solids. Phys. Rev. E—Stat. Nonlinear Soft Matter Phys. 2005, 71, 036123. [Google Scholar] [CrossRef]
  180. Ciavarella, M. A simplified version of Persson’s multiscale theory for rubber friction due to viscoelastic losses. J. Tribol. 2018, 140, 011403. [Google Scholar] [CrossRef]
  181. Persson, B.N. Rubber friction: Role of the flash temperature. J. Phys. Condens. Matter 2006, 18, 7789. [Google Scholar] [CrossRef]
  182. Carbone, G.; Persson, B. Crack motion in viscoelastic solids: The role of the flash temperature. Eur. Phys. J. E 2005, 17, 261–281. [Google Scholar] [CrossRef]
  183. Persson, B. Role of frictional heating in rubber friction. Tribol. Lett. 2014, 56, 77–92. [Google Scholar] [CrossRef]
  184. Palade, L.; Verney, V.; Attane, P. Time-temperature superposition and linear viscoelasticity of polybutadienes. Macromolecules 1995, 28, 7051–7057. [Google Scholar] [CrossRef]
  185. Schallamach, A. How does rubber slide? Wear 1971, 17, 301–312. [Google Scholar] [CrossRef]
  186. Tolpekina, T.; Pyckhout-Hintzen, W.; Persson, B. Linear and nonlinear viscoelastic modulus of rubber. Lubricants 2019, 7, 22. [Google Scholar] [CrossRef]
  187. Genovese, A.; Carputo, F.; Ciavarella, M.; Farroni, F.; Papangelo, A.; Sakhnevych, A. Analysis of multiscale theories for viscoelastic rubber friction. In Proceedings of the Conference of the Italian Association of Theoretical and Applied Mechanics, Rome, Italy, 15–19 September 2019; Springer: Cham, Switzerland, 2019; pp. 1125–1135. [Google Scholar]
  188. Fina, E.; Gruber, P.; Sharp, R. Hysteretic rubber friction: Application of Persson’s theories to Grosch’s experimental results. J. Appl. Mech. 2014, 81, 121001. [Google Scholar] [CrossRef]
  189. Furlan, M.; Mavros, G. A neural network approach for roughness-dependent update of tyre friction. Simul. Model. Pract. Theory 2022, 116, 102484. [Google Scholar] [CrossRef]
Figure 1. Elastic half-space indented by a rigid sphere of radius R [20].
Figure 1. Elastic half-space indented by a rigid sphere of radius R [20].
Applsci 15 11558 g001
Figure 2. Quadratic pressure distribution of Hertzian contact under different vertical loads.
Figure 2. Quadratic pressure distribution of Hertzian contact under different vertical loads.
Applsci 15 11558 g002
Figure 3. Contact between a rigid plane and a nominally flat but rough surface. Only asperities taller than the separation d contribute to load support [69].
Figure 3. Contact between a rigid plane and a nominally flat but rough surface. Only asperities taller than the separation d contribute to load support [69].
Applsci 15 11558 g003
Figure 4. Schematic representation of the Greenwood–Williamson statistical asperity model [69].
Figure 4. Schematic representation of the Greenwood–Williamson statistical asperity model [69].
Applsci 15 11558 g004
Figure 5. Comparison between JKR and Hertz theory for a rubber sphere in dry contact with a flat surface [86].
Figure 5. Comparison between JKR and Hertz theory for a rubber sphere in dry contact with a flat surface [86].
Applsci 15 11558 g005
Figure 6. Schematic comparison of the DMT and JKR adhesive contact models [25].
Figure 6. Schematic comparison of the DMT and JKR adhesive contact models [25].
Applsci 15 11558 g006
Figure 7. Rubber block sliding with velocity v on a rough counter surface under vertical load F.
Figure 7. Rubber block sliding with velocity v on a rough counter surface under vertical load F.
Applsci 15 11558 g007
Figure 8. Nominal contact area A 0 and contributions to the macroscopic real contact area A c , 0 .
Figure 8. Nominal contact area A 0 and contributions to the macroscopic real contact area A c , 0 .
Applsci 15 11558 g008
Figure 9. Ratio of real to nominal contact area as a function of sliding velocity at different pressures, with model results following Klüppel–Heinrich prediction computed and plotted by the authors [116].
Figure 9. Ratio of real to nominal contact area as a function of sliding velocity at different pressures, with model results following Klüppel–Heinrich prediction computed and plotted by the authors [116].
Applsci 15 11558 g009
Figure 10. Friction forces developed during tire–road interaction [155].
Figure 10. Friction forces developed during tire–road interaction [155].
Applsci 15 11558 g010
Figure 11. Fitted parameter b, computed and plotted by the authors (see [178]).
Figure 11. Fitted parameter b, computed and plotted by the authors (see [178]).
Applsci 15 11558 g011
Figure 12. Dry friction coefficients on coarse granite vs. sliding velocity for different pressures and temperatures; model from [116], replotted by the authors.
Figure 12. Dry friction coefficients on coarse granite vs. sliding velocity for different pressures and temperatures; model from [116], replotted by the authors.
Applsci 15 11558 g012
Table 1. Comparison of contact models.
Table 1. Comparison of contact models.
ModelsStrengthsLimitations
HertzSimple and analytically solvable
Applicable for smooth, elastic bodies
Closed-form solutions for stress, deformation and contact radius
Assumes idealized smooth surfaces
Ignores roughness and adhesion
Limited to small elastic deformations
JKRIncorporates adhesion into contact analysis
Relevant for soft materials and micro/nanoscale contacts
Extends Hertz to include surface energy effects
Assumes perfectly smooth surfaces
Limited for rough or multiscale contacts
Strong dependence on surface energy estimates
Greenwood–WilliamsonIntroduces statistical roughness description
Estimates contact area and pressure via asperity height distributions
Assumes identical, spherical, non-interacting asperities
Neglects elastic coupling and scale dependence
Klüppel and HeinrichIncludes viscoelastic effects and scaling
Defines minimum contact length and real contact evolution
Requires empirical cutoff parameters
Relies on simplified statistical descriptors
PerssonFully multiscale, PSD-based formulation
Accounts for roughness and viscoelasticity across scales
Mathematically complex
Sensitive to cutoff wavevectors and magnification
Table 2. Comparison of friction theories.
Table 2. Comparison of friction theories.
ModelsStrengthsLimitations
Grosch Tests (1963)First to demonstrate μ –frequency dependence
Linked friction to viscoelastic modulus peaks
Provided basis for later theoretical models
Purely empirical: no predictive capability
Cannot compute μ from compound or surface parameters
Klüppel and HeinrichSemi-analytical, multiscale approach
Evaluates hysteresis and adhesion contributions
Incorporates cutoff length and viscoelastic losses
Adhesion modeled empirically
Requires parameter fitting
No flash temperature treatment
PerssonFully multiscale framework using PSD
Directly links friction to viscoelastic moduli
Includes flash temperature effects
Computationally expensive
Sensitive to cutoff wavevector
Adhesion term remains empirical
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Stefanelli, R.; Fichera, G.; Genovese, A.; Napolitano Dell’Annunziata, G.; Sakhnevych, A.; Timpone, F.; Farroni, F. A Comprehensive Review of Rubber Contact Mechanics and Friction Theories. Appl. Sci. 2025, 15, 11558. https://doi.org/10.3390/app152111558

AMA Style

Stefanelli R, Fichera G, Genovese A, Napolitano Dell’Annunziata G, Sakhnevych A, Timpone F, Farroni F. A Comprehensive Review of Rubber Contact Mechanics and Friction Theories. Applied Sciences. 2025; 15(21):11558. https://doi.org/10.3390/app152111558

Chicago/Turabian Style

Stefanelli, Raffaele, Gabriele Fichera, Andrea Genovese, Guido Napolitano Dell’Annunziata, Aleksandr Sakhnevych, Francesco Timpone, and Flavio Farroni. 2025. "A Comprehensive Review of Rubber Contact Mechanics and Friction Theories" Applied Sciences 15, no. 21: 11558. https://doi.org/10.3390/app152111558

APA Style

Stefanelli, R., Fichera, G., Genovese, A., Napolitano Dell’Annunziata, G., Sakhnevych, A., Timpone, F., & Farroni, F. (2025). A Comprehensive Review of Rubber Contact Mechanics and Friction Theories. Applied Sciences, 15(21), 11558. https://doi.org/10.3390/app152111558

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop