Bulk and fracture process zone contribution to the rate-dependent adhesion amplification in viscoelastic broad-band materials

Ali Maghami Qingao Wang Michele Tricarico Michele Ciavarella Qunyang Li Antonio Papangelo
Abstract

The contact between a rigid Hertzian indenter and an adhesive broad-band viscoelastic substrate is considered. The material behaviour is described by a modified power law model, which is characterized by only four parameters, the glassy and rubbery elastic moduli, a characteristic exponent n𝑛nitalic_n and a timescale τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The maximum adherence force that can be reached while unloading the rigid indenter from a relaxed viscoelastic half-space is studied by means of a numerical implementation based on the boundary element method, as a function of the unloading velocity, preload and by varying the broadness of the viscoelastic material spectrum. Through a comprehensive numerical analysis we have determined the minimum contact radius that is needed to achieve the maximum amplification of the pull-off force at a specified unloading rate and for different material exponents n𝑛nitalic_n. The numerical results are then compared with the prediction of Persson and Brener viscoelastic crack propagation theory, providing excellent agreement. However, comparison against experimental tests for a glass lens indenting a PDMS substrate show data can be fitted with the linear theory only up to an unloading rate of about 100 μ100 μ100\textrm{ $\mu$}100 italic_μm/s showing the fracture process zone rate-dependent contribution to the energy enhancement is of the same order of the bulk dissipation contribution. Hence, the limitations of the current numerical and theoretical models for viscoelastic adhesion are discussed in light of the most recent literature results.

keywords:
Adhesion, Viscoelasticity, Sphere contact, Enhancement, Pull-off , Modified power law , Surface energy
journal: Elsevier
\affiliation

[inst1]organization=Politecnico di Bari, Department of Mechanics Mathematics and Management, addressline=Via Orabona 4, city=Bari, postcode=70125, country=Italy

\affiliation

[inst2]organization=AML, Department of Engineering Mechanics, Tsinghua University, city=Bei**g, postcode=100084, country=China

\affiliation

[inst3]organization=Hamburg University of Technology, Department of Mechanical Engineering,addressline=Am Schwarzenberg-Campus 1, city=Hamburg, postcode=21073, country=Germany

{graphicalabstract}[Uncaptioned image]
{highlights}

Numerical, theoretical and experimental investigations on viscoelastic adhesion enhancement

Maximum amplification is constrained by preload and unloading rate

Closed form results for the effective surface energy in broad-band spectrum viscoelastic materials

Bulk and fracture process zone contribution to the effective surface energy

Experiments show limitations of classical linear theories

1 Introduction

Understanding the adhesive behavior of soft materials such as polymers and elastomers would be of interest in many engineering applications, ranging from friction (Lorenz et al., 2015; Peng et al., 2021; Nazari et al., 2024; Mandriota et al., 2024), grip** technologies (Shintake et al., 2018), switchable adhesion (Kamperman et al., 2010; Papangelo and Ciavarella, 2017; Linghu et al., 2023, 2024), bio-mechanics (Felicetti et al., 2022; Forsbach et al., 2023), and soft robotics (Mazzolai et al., 2019; Agnelli et al., 2021). Since the fundamental contribution of Johnson, Kendall, and Roberts, the JKR contact model (Johnson et al., 1971), it is known that adhesion of compliant elastic materials can be understood as a Griffith energy balance between the elastic strain energy released as the crack advances and the energy dissipated by the formation of new surfaces, as it is classically known in Linear Elastic Fracture Mechanics (Maugis, 1992). In this case, for quasi-static conditions a thermodynamic work of adhesion Δγ0Δsubscript𝛾0\Delta\gamma_{0}roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, sometimes referred to as “surface energy”, can be defined that is independent of the rate at which the remote load is applied.

If the crack is advancing in a viscoelastic media, care should be taken in accounting for the dissipative effects introduced by viscoelasticity. Currently, few theories attempt a description of this problem. The cohesive zone models (Schapery, 1975a, b; Greenwood and Johnson, 1981; Greenwood, 2004; Schapery, 2022b) attempt an accurate description of the adhesive interactions taking place at the crack mouth, more precisely within a ”process zone” of length l0subscript𝑙0l_{0}italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT where the material bonds are actually disrupted and which is a function of the crack speed, resulting in the dependence on the speed of the effective surface energy ΔγeffΔsubscript𝛾𝑒𝑓𝑓\Delta\gamma_{eff}roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT. A different model developed by Persson and Brener (2005) (PB theory in the following) focuses on a steady-state moving crack. It establishes an energy balance between the power supplied by the applied load and the power dissipated by viscoelastic losses within the bulk material and the creation of new surfaces. By introducing a critical stress threshold for the rupture of material bonding, σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, it also introduces a typical lengthscale l0subscript𝑙0l_{0}italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT to calculate the dissipation in the bulk viscoelastic material (Persson and Brener, 2005; Persson, 2017, 2021).

Both cohesive and dissipation models really require a reference stress, or a reference length scale, which ultimately is used as a free parameter to fit the data points, and the linear theories result sometimes in nonphysical size for the fracture process zone at low speeds (Hui et al., 2022), as we shall discuss with reference to our results in the Discussion paragraph. Schapery developed also more elaborate theories (Schapery, 1984) to include non linear stress-strain behaviour using J𝐽Jitalic_J integral and also far field viscoelasticity, which require obviously even more material characterization.

For a semi-infinite system, the cohesive-model approach and PB theory lead to a monotonic increase of the effective (or ”apparent”) surface energy ΔγeffΔsubscript𝛾𝑒𝑓𝑓\Delta\gamma_{eff}roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT with the velocity v𝑣vitalic_v up to the theoretical “high-frequency” limit of Δγeff/Δγ0=E/E0Δsubscript𝛾𝑒𝑓𝑓Δsubscript𝛾0subscript𝐸subscript𝐸0\Delta\gamma_{eff}/\Delta\gamma_{0}=E_{\infty}/E_{0}roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT / roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT being Esubscript𝐸E_{\infty}italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT and E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT respectively the glassy and the rubbery elastic moduli of the viscoelastic material (Ciavarella et al., 2021). For the unloading of a flat-punch from a viscoelastic substrate, both approaches have been recently revisited to include finite-size effects, which have been showed to still give a monotonic increase of the effective surface energy, but with a maximum amplification that is limited by the system dimension providing max(Δγeff/Δγ0)<E/E0Δsubscript𝛾𝑒𝑓𝑓Δsubscript𝛾0subscript𝐸subscript𝐸0\max\left(\Delta\gamma_{eff}/\Delta\gamma_{0}\right)<E_{\infty}/E_{0}roman_max ( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT / roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) < italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (Maghami et al., 2024). Even for systems that can be considered semi-infinite, the contact problem presents several challenges as macroscopic adhesion is generally influenced by the indenter geometry (Papangelo and Ciavarella, 2023; Maghami et al., 2024) and the contact history (Greenwood and Johnson, 1981; Violano et al., 2021; Violano and Afferrante, 2022; Afferrante and Violano, 2022; VanDonselaar et al., 2023).

To precisely assess adhesion in soft polymers (silicone, rubber), the properties of the viscoelastic material need to be characterized. Several numerical and experimental works have tried to accurately determine the viscoelastic material response in the time domain (Wayne Chen et al., 2011; Lin et al., 2022; Dusane et al., 2023; Qi et al., 2024), in the frequency domain (Huang et al., 2004; Efremov et al., 2017) or using big data analysis and machine learning algorithms (Saharuddin et al., 2020; Hosseini et al., 2021). All the approaches reveal that real rubbers and elastomers have to be characterized over a very wide range of frequencies which typically spans many orders of magnitude of the exciting frequency (broad-band behavior), and this, in turn, plays a crucial role in determining the bulk dissipation, hence the interfacial adherence force.

Here we shall consider the problem of a rigid sphere of radius R𝑅Ritalic_R that is unloaded from a fully relaxed broad-band viscoelastic adhesive half-space (see Fig. 1) presenting and comparing numerical, analytical and experimental results. A few recent works (Violano et al., 2021, 2022; Afferrante and Violano, 2022) have focused on the problem of the adhesion of a rigid Hertzian indenter unloaded from a viscoelastic substrate describing the material either by using (i) the classical three-element solid, also known as the Standard Linear Solid ”SLS” (a spring in series with an element constituted by a dashpot and a spring in parallel), (Müser and Persson, 2022; Violano et al., 2022; Afferrante and Violano, 2022) or (ii) by considering the measured response spectrum of the viscoelastic material used in the experimental campaign (Violano et al., 2021, 2022; VanDonselaar et al., 2023). The limitation of the first approach is that the SLS has a narrow-band behavior, hence, although providing valuable insights, it will be rarely useful for modeling the behavior of a real material. The limitation of the second approach is that the results obtained solve only the specific problem considered and its difficult to draw general conclusions.

The objectives of this work are: (i) to define a material model which may be effectively and efficiently used for describing a real material viscoelastic behavior with a minimal number of constants in both the time and frequency domain, (ii) to determine, for the case of rigid Hertzian indenter-viscoelastic halfspace contact (see Fig. 1), how the broad-band material behavior influences the maximum adherence force as a function of the unloading rate (which is not the speed of the contact radius change, as we shall see) and of the preload, (iii) to provide closed form results for the effective surface energy ΔγeffΔsubscript𝛾𝑒𝑓𝑓\Delta\gamma_{eff}roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT based on the Persson and Brener (2005) theory which allows to faithfully reproduce the numerical results together with their region of validity, (iv) to validate the proposed approach by comparing the numerical results with experimental data.

Refer to caption
Figure 1: Sketch of the geometrical model: (a) indentation phase, (b) unloading phase with a constant unloading rate; (c) loading protocol consisting of pre-loading, dwelling, and unloading; (d) the Lennard-Jones force-separation law used at the interface.

The remainder of the paper is structured to address each objective outlined earlier: Section 2 provides a detailed description of the modified power law model used to characterize the viscoelastic material response; Section 3 presents, the Boundary Element Model developed for analyzing the adhesive contact problem along with extensive numerical results; Section 5 introduces the developed analytical solution and demonstrates its applicability in characterizing adhesive contact problems; Section 6 focuses on the validation of the numerical results based on experimental outcomes; Section 7 discusses the presented results in light of the recent Literature; the manuscript closes with the “Conclusions”, Section 8.

2 Modified power law model

The challenge behind the mechanical modeling of viscoelastic materials arises because the mechanical response at time t𝑡titalic_t depends on the contact history, so that the stress σ(t)𝜎𝑡\sigma\left(t\right)italic_σ ( italic_t ) and strain ε(t)𝜀𝑡\varepsilon\left(t\right)italic_ε ( italic_t ) should be determined by superposition:

ε(t)𝜀𝑡\displaystyle\varepsilon\left(t\right)italic_ε ( italic_t ) =σ(0)C(t)+0tC(tτ)dσ(τ)dτ𝑑τ,absent𝜎0𝐶𝑡superscriptsubscript0𝑡𝐶𝑡𝜏𝑑𝜎𝜏𝑑𝜏differential-d𝜏\displaystyle=\sigma\left(0\right)C\left(t\right)+\int_{0}^{t}C\left(t-\tau% \right)\frac{d\sigma\left(\tau\right)}{d\tau}d\tau\;,= italic_σ ( 0 ) italic_C ( italic_t ) + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_C ( italic_t - italic_τ ) divide start_ARG italic_d italic_σ ( italic_τ ) end_ARG start_ARG italic_d italic_τ end_ARG italic_d italic_τ , (1)
σ(t)𝜎𝑡\displaystyle\sigma\left(t\right)italic_σ ( italic_t ) =ε(0)R(t)+0tR(tτ)dε(τ)dτ𝑑τ,absent𝜀0𝑅𝑡superscriptsubscript0𝑡𝑅𝑡𝜏𝑑𝜀𝜏𝑑𝜏differential-d𝜏\displaystyle=\varepsilon\left(0\right)R\left(t\right)+\int_{0}^{t}R\left(t-% \tau\right)\frac{d\varepsilon\left(\tau\right)}{d\tau}d\tau\;,= italic_ε ( 0 ) italic_R ( italic_t ) + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_R ( italic_t - italic_τ ) divide start_ARG italic_d italic_ε ( italic_τ ) end_ARG start_ARG italic_d italic_τ end_ARG italic_d italic_τ , (2)

where C(t)𝐶𝑡C\left(t\right)italic_C ( italic_t ) is the creep compliance function, giving the strain response to a unit stress increment σ(t)𝜎𝑡\sigma\left(t\right)italic_σ ( italic_t ) in uniaxial loading conditions, and R(t)𝑅𝑡R\left(t\right)italic_R ( italic_t ) is the relaxation function giving the stress response to a unit strain increment ε(t)𝜀𝑡\varepsilon\left(t\right)italic_ε ( italic_t ) in uniaxial loading conditions. Alternatively, viscoelastic materials can be characterized in the frequency domain. If a sinusoidal stress σ(ω)𝜎𝜔\sigma\left(\omega\right)italic_σ ( italic_ω ) at frequency ω𝜔\omegaitalic_ω is applied to a viscoelastic specimen the resulting harmonic strain ε(ω)𝜀𝜔\varepsilon\left(\omega\right)italic_ε ( italic_ω ) will be delayed by a certain amount δ𝛿\deltaitalic_δ, hence the so-called complex modulus E¯(ω)=σ(ω)/ε(ω)¯𝐸𝜔𝜎𝜔𝜀𝜔\overline{E}\left(\omega\right)=\sigma\left(\omega\right)/\varepsilon\left(% \omega\right)\ over¯ start_ARG italic_E end_ARG ( italic_ω ) = italic_σ ( italic_ω ) / italic_ε ( italic_ω )can be defined in the complex plane. Alternatively, in place of E¯(ω)¯𝐸𝜔\overline{E}\left(\omega\right)over¯ start_ARG italic_E end_ARG ( italic_ω ) one may define its reciprocal, the C¯(ω)=ε(ω)/σ(ω)¯𝐶𝜔𝜀𝜔𝜎𝜔\overline{C}\left(\omega\right)=\varepsilon\left(\omega\right)/\sigma\left(% \omega\right)over¯ start_ARG italic_C end_ARG ( italic_ω ) = italic_ε ( italic_ω ) / italic_σ ( italic_ω ) which is the complex compliance.

One approach to reproduce the broad-band response spectrum of a real viscoelastic material (VanDonselaar et al., 2023; Lorenz et al., 2013) is to move from a SLS material model, which is constituted by a spring in parallel with a single Maxwell element (a spring in series with a dashpot), to the so-called Wiechert model constituted by a spring in parallel with many Maxwell elements (Christensen, 2012), so that several relaxation times can be included in the material representation. Very often the number of elements needed for a faithful representation gets large enough so that the model returns a very good representation of the material viscoelastic behaviour, but at the same time it makes it difficult to extract general conclusions, due to the large number of fitting parameters determined.

One option to overcome this difficulty is to rely on power law material models (Schapery, 1975a; Persson and Brener, 2005; Popov et al., 2010; Bonfanti et al., 2020; Dusane et al., 2023), which assume a certain power law function for the distribution of the relaxation times. For example Popov et al. (2010) proposes a model that is fully defined by 5 constants: the relaxed and glassy moduli, two characteristic times, and one exponent. Schapery (1975a) uses an approximation for the creep compliance function C(t)=(Me+M1tp)1𝐶𝑡superscriptsubscript𝑀𝑒subscript𝑀1superscript𝑡𝑝1C\left(t\right)=\left(M_{e}+M_{1}t^{-p}\right)^{-1}italic_C ( italic_t ) = ( italic_M start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT + italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_t start_POSTSUPERSCRIPT - italic_p end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT which includes only three material constants {Me,M1,p}subscript𝑀𝑒subscript𝑀1𝑝\left\{M_{e},M_{1},p\right\}{ italic_M start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT , italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_p } and can describe well the behavior for very long times while being less accurate in describing the short-time material behavior. Furthermore, Persson and Brener (2005) consider a model where the retardation times are distributed as a power law in between two characteristic times and vanishes elsewhere.

In the following, we will consider and extend the Modified Power Law (MPL) material model introduced by Williams (1964), which is fully defined by a minimal set of four parameters: the glassy Esubscript𝐸E_{\infty}italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT and the rubbery E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT moduli, a single characteristic time τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and one exponent n𝑛nitalic_n. Closed-form results in both time and frequency domain that can be readily used for real viscoelastic material characterization or as input in viscoelastic crack propagation theories are provided in A, while in the following the main results are reported.

For the MPL material, the following relaxation spectrum H(τ)𝐻𝜏H(\tau)italic_H ( italic_τ ) is assumed

H(τ)=(EE0Γ(n))(τ0τ)nexp(τ0τ),𝐻𝜏subscript𝐸subscript𝐸0Γ𝑛superscriptsubscript𝜏0𝜏𝑛subscript𝜏0𝜏H\left(\tau\right)=\left(\frac{E_{\infty}-E_{0}}{\Gamma\left(n\right)}\right)% \left(\frac{\tau_{0}}{\tau}\right)^{n}\exp\left(-\frac{\tau_{0}}{\tau}\right)\;,italic_H ( italic_τ ) = ( divide start_ARG italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG roman_Γ ( italic_n ) end_ARG ) ( divide start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_τ end_ARG ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT roman_exp ( - divide start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_τ end_ARG ) , (3)

where {τ0,n}subscript𝜏0𝑛\left\{\tau_{0},n\right\}{ italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_n } are constants to be determined and Γ(n)Γ𝑛\Gamma(n)roman_Γ ( italic_n ) is the Gamma function. The substitution of Eq. (3) into Eq. (33) gives the complex modulus E¯(ω)=E(ω)+𝒊E′′(ω)¯𝐸𝜔superscript𝐸𝜔𝒊superscript𝐸′′𝜔\overline{E}\left(\omega\right)=E^{\prime}\left(\omega\right)+\boldsymbol{i}E^% {\prime\prime}\left(\omega\right)over¯ start_ARG italic_E end_ARG ( italic_ω ) = italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω ) + bold_italic_i italic_E start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ω ):

E¯(ω)=E0+(EE0)𝒊ωτ0exp(𝒊ωτ0)𝐄n(𝒊ωτ0),¯𝐸𝜔subscript𝐸0subscript𝐸subscript𝐸0𝒊𝜔subscript𝜏0𝒊𝜔subscript𝜏0subscript𝐄𝑛𝒊𝜔subscript𝜏0\overline{E}\left(\omega\right)=E_{0}+\left(E_{\infty}-E_{0}\right)\boldsymbol% {i}\omega\tau_{0}\exp\left(\boldsymbol{i}\omega\tau_{0}\right)\mathbf{E}_{n}% \left(\boldsymbol{i}\omega\tau_{0}\right)\;,over¯ start_ARG italic_E end_ARG ( italic_ω ) = italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + ( italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) bold_italic_i italic_ω italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_exp ( bold_italic_i italic_ω italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) bold_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_italic_i italic_ω italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) , (4)

where 𝐄n(x)subscript𝐄𝑛𝑥\mathbf{E}_{n}\left(x\right)bold_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) is the exponential integral function of order n>0𝑛0n>0italic_n > 0. A provides closed-form results for both the real E(ω)superscript𝐸𝜔E^{\prime}\left(\omega\right)italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω ) and the imaginary parts E′′(ω)superscript𝐸′′𝜔E^{\prime\prime}\left(\omega\right)italic_E start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ω ) of the complex modulus.

The relaxation function R(t)𝑅𝑡R\left(t\right)italic_R ( italic_t ) is given by (Williams, 1964):

R(t)=E0+0τ1H(τ)exp(t/τ)𝑑τ,𝑅𝑡subscript𝐸0superscriptsubscript0superscript𝜏1𝐻𝜏𝑡𝜏differential-d𝜏R\left(t\right)=E_{0}+\int_{0}^{\infty}\tau^{-1}H\left(\tau\right)\exp\left(-t% /\tau\right)d\tau\;,italic_R ( italic_t ) = italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_H ( italic_τ ) roman_exp ( - italic_t / italic_τ ) italic_d italic_τ , (5)

which, upon substitution of Eq. (3) gives a very simple form:

R(t)=E0+EE0(1+t/τ0)n𝑅𝑡subscript𝐸0subscript𝐸subscript𝐸0superscript1𝑡subscript𝜏0𝑛R\left(t\right)=E_{0}+\frac{E_{\infty}-E_{0}}{\left(1+t/\tau_{0}\right)^{n}}\;italic_R ( italic_t ) = italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + divide start_ARG italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG ( 1 + italic_t / italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG (6)

or in dimensionless form

R^(t^)=1+1/k1(1+t^)n^𝑅^𝑡11𝑘1superscript1^𝑡𝑛\widehat{R}\left(\widehat{t}\right)=1+\frac{1/k-1}{\left(1+\widehat{t}\right)^% {n}}\;over^ start_ARG italic_R end_ARG ( over^ start_ARG italic_t end_ARG ) = 1 + divide start_ARG 1 / italic_k - 1 end_ARG start_ARG ( 1 + over^ start_ARG italic_t end_ARG ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG (7)

being R^=R(t)/E0^𝑅𝑅𝑡subscript𝐸0\widehat{R}=R\left(t\right)/E_{0}over^ start_ARG italic_R end_ARG = italic_R ( italic_t ) / italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, t^=t/τ0^𝑡𝑡subscript𝜏0\widehat{t}=t/\tau_{0}over^ start_ARG italic_t end_ARG = italic_t / italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and k=E0/E𝑘subscript𝐸0subscript𝐸k=E_{0}/E_{\infty}italic_k = italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT, which shows that at a given dimensionless time t^^𝑡\widehat{t}over^ start_ARG italic_t end_ARG the material relaxation depends only on the parameters {n,k}𝑛𝑘\{n,k\}{ italic_n , italic_k }.

Similarly, we can assume a modified power law distribution for the retardation spectrum

L(τ)=(C0CΓ(n))(ττ0)nexp(ττ0),𝐿𝜏subscript𝐶0subscript𝐶Γ𝑛superscript𝜏subscript𝜏0𝑛𝜏subscript𝜏0L\left(\tau\right)=\left(\frac{C_{0}-C_{\infty}}{\Gamma\left(n\right)}\right)% \left(\frac{\tau}{\tau_{0}}\right)^{n}\exp\left(-\frac{\tau}{\tau_{0}}\right)\;,italic_L ( italic_τ ) = ( divide start_ARG italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG start_ARG roman_Γ ( italic_n ) end_ARG ) ( divide start_ARG italic_τ end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT roman_exp ( - divide start_ARG italic_τ end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) , (8)

where {τ0,n}subscript𝜏0𝑛\left\{\tau_{0},n\right\}{ italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_n } are constants to be determined. Hence the complex compliance is (substitute Eq. (8) into Eq. (47)):

C¯(ω)=C+(C0C)𝒊ωτ0exp(𝒊ωτ0)𝐄n(𝒊ωτ0),¯𝐶𝜔subscript𝐶subscript𝐶0subscript𝐶𝒊𝜔subscript𝜏0𝒊𝜔subscript𝜏0subscript𝐄𝑛𝒊𝜔subscript𝜏0\overline{C}\left(\omega\right)=C_{\infty}+\frac{\left(C_{0}-C_{\infty}\right)% }{\boldsymbol{i}\omega\tau_{0}}\exp\left(-\frac{\boldsymbol{i}}{\omega\tau_{0}% }\right)\mathbf{E}_{n}\left(-\frac{\boldsymbol{i}}{\omega\tau_{0}}\right)\;,over¯ start_ARG italic_C end_ARG ( italic_ω ) = italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT + divide start_ARG ( italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ) end_ARG start_ARG bold_italic_i italic_ω italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG roman_exp ( - divide start_ARG bold_italic_i end_ARG start_ARG italic_ω italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) bold_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( - divide start_ARG bold_italic_i end_ARG start_ARG italic_ω italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) , (9)

where 𝐄n(x)subscript𝐄𝑛𝑥\mathbf{E}_{n}\left({x}\right)bold_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) is the exponential integral function of order n>0𝑛0n>0italic_n > 0, C0=1/E0subscript𝐶01subscript𝐸0C_{0}=1/E_{0}italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 / italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the creep compliance in the rubbery limit and C=1/Esubscript𝐶1subscript𝐸C_{\infty}=1/E_{\infty}italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 1 / italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT is the creep compliance in the glassy limit. Notice that once C¯(ω)¯𝐶𝜔\overline{C}\left(\omega\right)over¯ start_ARG italic_C end_ARG ( italic_ω ) is obtained, the complex modulus is also obtained as E¯(ω)=1/C¯(ω)¯𝐸𝜔1¯𝐶𝜔\overline{E}\left(\omega\right)=1/\overline{C}\left(\omega\right)over¯ start_ARG italic_E end_ARG ( italic_ω ) = 1 / over¯ start_ARG italic_C end_ARG ( italic_ω ) and vice-versa. The A reports closed form results for both the real C(ω)superscript𝐶𝜔C^{\prime}\left(\omega\right)italic_C start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω ) and the imaginary part C′′(ω)superscript𝐶′′𝜔C^{\prime\prime}\left(\omega\right)italic_C start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ω ) of C¯(ω)¯𝐶𝜔\overline{C}\left(\omega\right)over¯ start_ARG italic_C end_ARG ( italic_ω ).

The creep compliance function C(t)𝐶𝑡C\left(t\right)italic_C ( italic_t ) is given by (Williams, 1964):

C(t)=C+0τ1L(τ)(1exp(t/τ))𝑑τ,𝐶𝑡subscript𝐶superscriptsubscript0superscript𝜏1𝐿𝜏1𝑡𝜏differential-d𝜏C\left(t\right)=C_{\infty}+\int_{0}^{\infty}\tau^{-1}L\left(\tau\right)\left(1% -\exp\left(-t/\tau\right)\right)d\tau\;,italic_C ( italic_t ) = italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_L ( italic_τ ) ( 1 - roman_exp ( - italic_t / italic_τ ) ) italic_d italic_τ , (10)

which, upon substitution of Eq. (8) into Eq. (10) gives:

C(t)=C02(C0C)Γ(n)(tτ0)n/2𝐊n(2tτ0),𝐶𝑡subscript𝐶02subscript𝐶0subscript𝐶Γ𝑛superscript𝑡subscript𝜏0𝑛2subscript𝐊𝑛2𝑡subscript𝜏0C\left(t\right)=C_{0}-2\frac{\left(C_{0}-C_{\infty}\right)}{\Gamma\left(n% \right)}\left(\frac{t}{\tau_{0}}\right)^{n/2}\mathbf{K}_{n}\left(2\sqrt{\frac{% t}{\tau_{0}}}\right)\;,italic_C ( italic_t ) = italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 2 divide start_ARG ( italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ) end_ARG start_ARG roman_Γ ( italic_n ) end_ARG ( divide start_ARG italic_t end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_n / 2 end_POSTSUPERSCRIPT bold_K start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( 2 square-root start_ARG divide start_ARG italic_t end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG end_ARG ) , (11)

where 𝐊n(x)subscript𝐊𝑛𝑥\mathbf{K}_{n}\left(x\right)bold_K start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) is the modified Bessel function of the second kind. The dimensionless creep compliance function C^=C/C0^𝐶𝐶subscript𝐶0\widehat{C}=C/C_{0}over^ start_ARG italic_C end_ARG = italic_C / italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is

C^(t^)=12(1k)Γ(n)t^n/2𝐊n(2t^),^𝐶^𝑡121𝑘Γ𝑛superscript^𝑡𝑛2subscript𝐊𝑛2^𝑡\widehat{C}\left(\widehat{t}\right)=1-2\frac{\left(1-k\right)}{\Gamma\left(n% \right)}\widehat{t}^{n/2}\mathbf{K}_{n}\left(2\sqrt{\widehat{t}}\right)\;,over^ start_ARG italic_C end_ARG ( over^ start_ARG italic_t end_ARG ) = 1 - 2 divide start_ARG ( 1 - italic_k ) end_ARG start_ARG roman_Γ ( italic_n ) end_ARG over^ start_ARG italic_t end_ARG start_POSTSUPERSCRIPT italic_n / 2 end_POSTSUPERSCRIPT bold_K start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( 2 square-root start_ARG over^ start_ARG italic_t end_ARG end_ARG ) , (12)

which shows that at a given dimensionless time t^^𝑡\widehat{t}over^ start_ARG italic_t end_ARG the material creep depends only on the parameters {n,k}𝑛𝑘\{n,k\}{ italic_n , italic_k }.

In Section 6, we will show that Polydimethylsiloxane (PDMS, 10:1 resin to curing agent ratio) one of the most common silicone-based polymer used in soft contact mechanics (Shintake et al., 2018; Sahli et al., 2019; Oliver et al., 2023) has a characteristic exponent at room temperature of n0.22similar-to-or-equals𝑛0.22n\simeq 0.22italic_n ≃ 0.22, which is close to what Williams (1964) found for unfilled HC rubber. In Fig. 2, we illustrate the time evolution of the relaxation and creep compliance functions of MPL for different exponents n𝑛nitalic_n (solid black lines) alongside with a comparison of the SLS viscoelastic behaviour (blue dashed lines). Fig. 2 shows that in order to obtain a behavior close to a standard material, we should set n1.6𝑛1.6n\approx 1.6italic_n ≈ 1.6, which implies PDMS has a much broader spectrum with respect to the SLS. It is recalled that for a SLS the dimensionless creep compliance function is C^(t^)=[1(k1)exp(t^)]^𝐶^𝑡delimited-[]1𝑘1^𝑡\widehat{C}(\widehat{t})=[1-(k-1)\exp{(-\widehat{t})}]over^ start_ARG italic_C end_ARG ( over^ start_ARG italic_t end_ARG ) = [ 1 - ( italic_k - 1 ) roman_exp ( - over^ start_ARG italic_t end_ARG ) ].

Refer to caption
Figure 2: Time evolution of (a) the relaxation function R(t)𝑅𝑡R\left(t\right)italic_R ( italic_t ) (Eq. 6), (b) the creep compliance function C(t)𝐶𝑡C(t)italic_C ( italic_t ) (Eq. 11) for n=[0.2,0.4,0.6]𝑛0.20.40.6n=[0.2,0.4,0.6]italic_n = [ 0.2 , 0.4 , 0.6 ] (solid black line) and a comparison with the behaviour of a SLS viscoelastic material (dashed blue line).

3 The numerical model

Let us consider the problem of a rigid sphere of radius R𝑅Ritalic_R that is unloaded from a fully relaxed viscoelastic adhesive half-space (see Fig. 1). To model the adhesive contact problem a numerical scheme based on the Boundary Element Method was implemented in the software MATLAB, together with a time marching algorithm, which follows the implementation by Papangelo and Ciavarella (2020, 2023). In the numerical model it is assumed that the interaction between the sphere and the substrate is governed by a Lennard-Jones force-separation law 111Strictly speaking Eq. (13) would hold for infinite parallel planes, nevertheless in adhesive contact mechanics it is often assumed that Eq. (13) holds also for slightly inclined surfaces, which is the so-called ”Derjaguin approximation”, see (Greenwood, 1997).:

σ(h)=8Δγ03h0[(h0h)3(h0h)9],𝜎8Δsubscript𝛾03subscript0delimited-[]superscriptsubscript03superscriptsubscript09\sigma\left(h\right)=-\frac{8\Delta\gamma_{0}}{3h_{0}}\left[\left(\frac{h_{0}}% {h}\right)^{3}-\left(\frac{h_{0}}{h}\right)^{9}\right]\;,italic_σ ( italic_h ) = - divide start_ARG 8 roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 3 italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG [ ( divide start_ARG italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_h end_ARG ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - ( divide start_ARG italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_h end_ARG ) start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT ] , (13)

where σ𝜎\sigmaitalic_σ is the interfacial stress (σ>0,𝜎0\sigma>0,italic_σ > 0 , when compressive), hhitalic_h the local gap, h0subscript0h_{0}italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT the equilibrium distance with the surface energy Δγ0=9316σ0h0Δsubscript𝛾09316subscript𝜎0subscript0\Delta\gamma_{0}=\frac{9\sqrt{3}}{16}\sigma_{0}h_{0}roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = divide start_ARG 9 square-root start_ARG 3 end_ARG end_ARG start_ARG 16 end_ARG italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The gap function is then written as:

h(r,t)=δ+h0+r22R+uz(r,t),𝑟𝑡𝛿subscript0superscript𝑟22𝑅subscript𝑢𝑧𝑟𝑡h(r,t)=-\delta+h_{0}+\frac{r^{2}}{2R}+u_{z}\left(r,t\right)\;,italic_h ( italic_r , italic_t ) = - italic_δ + italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + divide start_ARG italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_R end_ARG + italic_u start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_r , italic_t ) , (14)

where δ>0𝛿0\delta>0italic_δ > 0 when the sphere approaches the viscoelastic half-space, the sphere profile is approximated by a parabola, and uz(r,t)subscript𝑢𝑧𝑟𝑡u_{z}\left(r,t\right)italic_u start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_r , italic_t ) is the deflection of the viscoelastic half-space, which depends on the loading history (we have explicitly shown the dependence of uzsubscript𝑢𝑧u_{z}italic_u start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT on time t𝑡titalic_t). The vertical deflections of the halfspace for an elastic axisymmetric problem are obtained as (Greenwood, 1997; Feng, 2000):

uz(r)=1Epsσ(s)G(r,s)s𝑑s,subscript𝑢𝑧𝑟1subscript𝐸𝑝𝑠𝜎𝑠𝐺𝑟𝑠𝑠differential-d𝑠u_{z}\left(r\right)=\frac{1}{E_{ps}}{\displaystyle\int}\sigma\left(s\right)G% \left(r,s\right)sds\;,italic_u start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_r ) = divide start_ARG 1 end_ARG start_ARG italic_E start_POSTSUBSCRIPT italic_p italic_s end_POSTSUBSCRIPT end_ARG ∫ italic_σ ( italic_s ) italic_G ( italic_r , italic_s ) italic_s italic_d italic_s , (15)

where G(r,s)𝐺𝑟𝑠G\left(r,s\right)italic_G ( italic_r , italic_s ) is the Kernel function:

G(r,s)={4πrK(sr),s<r4πsK(rs),s>r𝐺𝑟𝑠cases4𝜋𝑟𝐾𝑠𝑟𝑠𝑟4𝜋𝑠𝐾𝑟𝑠𝑠𝑟G\left(r,s\right)=\left\{\begin{array}[c]{cc}\frac{4}{\pi r}K\left(\frac{s}{r}% \right),&\qquad\qquad s<r\\ \frac{4}{\pi s}K\left(\frac{r}{s}\right),&\qquad\qquad s>r\end{array}\right.italic_G ( italic_r , italic_s ) = { start_ARRAY start_ROW start_CELL divide start_ARG 4 end_ARG start_ARG italic_π italic_r end_ARG italic_K ( divide start_ARG italic_s end_ARG start_ARG italic_r end_ARG ) , end_CELL start_CELL italic_s < italic_r end_CELL end_ROW start_ROW start_CELL divide start_ARG 4 end_ARG start_ARG italic_π italic_s end_ARG italic_K ( divide start_ARG italic_r end_ARG start_ARG italic_s end_ARG ) , end_CELL start_CELL italic_s > italic_r end_CELL end_ROW end_ARRAY (16)

and K(k)𝐾𝑘K\left(k\right)italic_K ( italic_k ) is the complete elliptic integral of the first kind of modulus k𝑘kitalic_k. Hence, according to the elastic-viscoelastic correspondence principle in the form of Boltzmann integrals (Christensen, 2012), the normal displacements of the viscoelastic half-space uz(r,t)subscript𝑢𝑧𝑟𝑡u_{z}(r,t)italic_u start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_r , italic_t ) at time t𝑡titalic_t, at position r𝑟ritalic_r will depend on the contact history as:

uz(r,t)=G(r,s)stC(tτ)dσ(s,τ)dτ𝑑τ𝑑s.subscript𝑢𝑧𝑟𝑡𝐺𝑟𝑠𝑠superscriptsubscript𝑡𝐶𝑡𝜏𝑑𝜎𝑠𝜏𝑑𝜏differential-d𝜏differential-d𝑠{u_{z}(r,t)=\int G\left(r,s\right)s\int_{-\infty}^{t}C(t-\tau)\frac{d\sigma(s,% \tau)}{d\tau}d\tau ds}\;.italic_u start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_r , italic_t ) = ∫ italic_G ( italic_r , italic_s ) italic_s ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_C ( italic_t - italic_τ ) divide start_ARG italic_d italic_σ ( italic_s , italic_τ ) end_ARG start_ARG italic_d italic_τ end_ARG italic_d italic_τ italic_d italic_s . (17)

where the moduli in the creep compliance function C(t)𝐶𝑡C(t)italic_C ( italic_t ) should be consider in the plane strain conditions, i.e. C0=1/E0=1ν2E0subscriptsuperscript𝐶01subscriptsuperscript𝐸01superscript𝜈2subscript𝐸0C^{*}_{0}=1/E^{*}_{0}=\frac{1-\nu^{2}}{E_{0}}italic_C start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 / italic_E start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = divide start_ARG 1 - italic_ν start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG, being ν𝜈\nuitalic_ν the Poisson ratio considered independent on the excitation frequency ω𝜔\omegaitalic_ω. The gap function Eq. (14) is solved through a Newton-Raphson scheme on N=M+1𝑁𝑀1N=M+1italic_N = italic_M + 1 equally-spaced nodes, being M𝑀Mitalic_M the number of interfacial elements so that Eq.s (13,14,17) are satisfied. To determine the half-space deflections Eq. (17) was discretized in time and space. In time, we used a time marching algorithm with a time step ΔtΔ𝑡\Delta troman_Δ italic_t. In space, we assumed the pressure distribution has a triangular shape over each element, i.e. for the element j𝑗jitalic_j-th the pressure is pjsubscript𝑝𝑗p_{j}italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT at r=rj𝑟subscript𝑟𝑗r=r_{j}italic_r = italic_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT and falls linearly to 00 at r=rj1𝑟subscript𝑟𝑗1r=r_{j-1}italic_r = italic_r start_POSTSUBSCRIPT italic_j - 1 end_POSTSUBSCRIPT and r=rj+1𝑟subscript𝑟𝑗1r=r_{j+1}italic_r = italic_r start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT, which is usually referred as the “method of the overlap** triangles” (Johnson, 1987). Further details of the numerical implementation can be found in Ref.s (Papangelo and Ciavarella, 2020, 2023).

Refer to caption
Refer to caption
Refer to caption
Figure 3: Dimensionless load P^^𝑃\widehat{P}over^ start_ARG italic_P end_ARG versus the dimensionless contact radius a^^𝑎\widehat{a}over^ start_ARG italic_a end_ARG. (a) Material exponent n=0.8𝑛0.8n=0.8italic_n = 0.8, initial contact radius a^0=3.61subscript^𝑎03.61\widehat{a}_{0}=3.61over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 3.61, unloading rates r^=[102,102.5,103,106]^𝑟superscript102superscript102.5superscript103superscript106\widehat{r}=[10^{2},10^{2.5},10^{3},10^{6}]over^ start_ARG italic_r end_ARG = [ 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 2.5 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ]; (b) initial contact radius a^0=3.61subscript^𝑎03.61\widehat{a}_{0}=3.61over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 3.61, unloading rate of r^=102.5^𝑟superscript102.5\widehat{r}=10^{2.5}over^ start_ARG italic_r end_ARG = 10 start_POSTSUPERSCRIPT 2.5 end_POSTSUPERSCRIPT for different material exponents n=[0.6,0.8,1.6]𝑛0.60.81.6n=[0.6,0.8,1.6]italic_n = [ 0.6 , 0.8 , 1.6 ]; (c) initial contact radii a^0=[1.71, 2.47, 3.61]subscript^𝑎01.712.473.61\widehat{a}_{0}=[1.71,\;2.47,\;3.61]over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = [ 1.71 , 2.47 , 3.61 ], with a constant unloading rate of r^=102.5^𝑟superscript102.5\widehat{r}=10^{2.5}over^ start_ARG italic_r end_ARG = 10 start_POSTSUPERSCRIPT 2.5 end_POSTSUPERSCRIPT and material exponent n=1.6𝑛1.6n=1.6italic_n = 1.6. For all the panels unloading starts from a fully relaxed substrate with k=0.1𝑘0.1k=0.1italic_k = 0.1 and μ=3.24𝜇3.24\mu=3.24italic_μ = 3.24:

4 Numerical results

In the rest of the paper, unless differently stated, the numerical results will be presented in dimensionless notation, as follows:

δ^=δ(π2Δγ02R/E02)1/3;a^=a(πR2Δγ0/E0)1/3;P^=PπΔγ0R,\begin{split}\widehat{\delta}=\frac{\delta}{(\pi^{2}\Delta\gamma_{0}^{2}R/{E_{% 0}^{*}}^{2})^{1/3}}\;;\quad\widehat{a}=\frac{a}{(\pi R^{2}\Delta\gamma_{0}/E_{% 0}^{*})^{1/3}}\;;\quad\widehat{P}=\frac{P}{\pi\Delta\gamma_{0}R}\;,\end{split}start_ROW start_CELL over^ start_ARG italic_δ end_ARG = divide start_ARG italic_δ end_ARG start_ARG ( italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_R / italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT end_ARG ; over^ start_ARG italic_a end_ARG = divide start_ARG italic_a end_ARG start_ARG ( italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT end_ARG ; over^ start_ARG italic_P end_ARG = divide start_ARG italic_P end_ARG start_ARG italic_π roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_R end_ARG , end_CELL end_ROW (18)

being δ^^𝛿\widehat{\delta}over^ start_ARG italic_δ end_ARG the dimensionless indentation, a^^𝑎\widehat{a}over^ start_ARG italic_a end_ARG the dimensionless contact radius, P^^𝑃\widehat{P}over^ start_ARG italic_P end_ARG the dimensionless normal load, P^po=|min(P^)|subscript^𝑃𝑝𝑜^𝑃\widehat{P}_{po}=|\min\left(\widehat{P}\right)|over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_p italic_o end_POSTSUBSCRIPT = | roman_min ( over^ start_ARG italic_P end_ARG ) | the maximum detachment force, i.e. the pull-off force. Unless specified otherwise, our simulations employ N=500𝑁500N=500italic_N = 500 nodes. Let’s consider a sphere with an initial contact radius of a0subscript𝑎0{a}_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is unloaded from a fully relaxed viscoelastic substrate, with various unloading rate r𝑟{r}italic_r. This unloading process mimics experimental conditions: (i) indenting the viscoelastic substrate to a specified depth (δloadsubscript𝛿load{\delta}_{\textrm{load}}italic_δ start_POSTSUBSCRIPT load end_POSTSUBSCRIPT), (ii) allowing dwell time for substrate relaxation, then (iii) unloading at a constant velocity r𝑟{r}italic_r. The corresponding dimensionless unloading rate is defined as r^=rτ/h0^𝑟𝑟𝜏subscript0\widehat{r}=r\tau/h_{0}over^ start_ARG italic_r end_ARG = italic_r italic_τ / italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Unless differently stated, the results provided in the following will refer to the Tabor parameter (Tabor, 1977) μ=(RΔγ02E02h03)1/3=3.24𝜇superscript𝑅Δsuperscriptsubscript𝛾02superscriptsubscriptsuperscript𝐸02superscriptsubscript03133.24\mu=\left(\frac{R{\Delta\gamma_{0}}^{2}}{{E^{*}_{0}}^{2}{h_{0}}^{3}}\right)^{1% /3}=3.24italic_μ = ( divide start_ARG italic_R roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_E start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT = 3.24 and k=E0/E=0.1𝑘subscript𝐸0subscript𝐸0.1k=E_{0}/E_{\infty}=0.1italic_k = italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 0.1.

4.1 Dependence of the detachment force upon the loading protocol details

As it was discussed, the unloading rate r𝑟ritalic_r has a crucial role in the mechanical response of viscoelastic materials. We examined our model for different unloading rates while Fig. 3 (a) reports only four different unloading rates of r^=[102,102.5,103,106]^𝑟superscript102superscript102.5superscript103superscript106\widehat{r}=[10^{2},10^{2.5},10^{3},10^{6}]over^ start_ARG italic_r end_ARG = [ 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 2.5 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ]. The sphere is unloaded from a fully relaxed substrate with the exponent material of n=0.8𝑛0.8n=0.8italic_n = 0.8 and all the unloading curves in Fig. 3 start from the initial contact radius of a^0=3.61subscript^𝑎03.61\widehat{a}_{0}=3.61over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 3.61. As anticipated in viscoelastic contact problems, the unloading rate significantly affects the unloading trajectory. Fast unloading boost viscoelastic dissipation at the crack tip which in turns gives a high pull-off load at detachment. Note that the elastic behavior observed in Fig. 3 corresponds to the initial state of the relaxed substrate.

For the same unloading rate r^=102.5^𝑟superscript102.5\widehat{r}=10^{2.5}over^ start_ARG italic_r end_ARG = 10 start_POSTSUPERSCRIPT 2.5 end_POSTSUPERSCRIPT, and starting from the same initial contact radius a^0=3.61subscript^𝑎03.61\widehat{a}_{0}=3.61over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 3.61 the unloading trajectory will be influenced by the response spectrum of the material. In particular, by using the MPL formulation for simulating a broad-band material (see Fig. 2), Fig. 3 (b) shows the unloading curves for n=[0.6,0.8,1.6]𝑛0.60.81.6n=[0.6,0.8,1.6]italic_n = [ 0.6 , 0.8 , 1.6 ], showing that for a given unloading rate the pull-off force generally increases by increasing n𝑛nitalic_n. As we will show later, this happens because the more narrow-band is the material response spectrum the faster the theoretical amplification Δγeff/Δγ0=E/E0Δsubscript𝛾𝑒𝑓𝑓Δsubscript𝛾0subscript𝐸subscript𝐸0\Delta\gamma_{eff}/\Delta\gamma_{0}=E_{\infty}/E_{0}roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT / roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is reached as a function of the retraction rate.

Refer to caption
Figure 4: Normalized pull-off force as a function of the normalized crack velocity for a material with power law exponent n=0.8𝑛0.8n=0.8italic_n = 0.8 and starting the unloading phase from a fully relaxed substrate with initial contact radii a^0=[1.94,2.50,3.61,5.28]subscript^𝑎01.942.503.615.28\widehat{a}_{0}=[1.94,2.50,3.61,5.28]over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = [ 1.94 , 2.50 , 3.61 , 5.28 ].

Finally, the maximum adhesion force at detachment is significantly influenced by the the preload, as it was discussed in (Violano and Afferrante, 2022; Afferrante and Violano, 2022). Changing the preload will affect the initial contact area, and also the initial indentation prior to unloading. To demonstrate this effect Fig. 3 (c) shows the unloading trajectories of three different unloading curves from the same relaxed viscoelastic substrate (here n=1.6𝑛1.6n=1.6italic_n = 1.6 and r^=102.5^𝑟superscript102.5\widehat{r}=10^{2.5}over^ start_ARG italic_r end_ARG = 10 start_POSTSUPERSCRIPT 2.5 end_POSTSUPERSCRIPT) while using a^0=[1.71,2.47,3.61]subscript^𝑎01.712.473.61\widehat{a}_{0}=[1.71,2.47,3.61]over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = [ 1.71 , 2.47 , 3.61 ] respectively. The larger is the initial contact radius the larger will be the pull-off force. Too small initial contact radius will give raise to finite size effects which will limit the possibility to enhance the pull-off force up to the theoretical limit predicted by viscoelastic crack propagation theories (Persson and Brener, 2005; Schapery, 1975a).

Hence, when looking for the maximum amplification of the pull-off force, care should be taken in selecting a large enough initial contact radius (or preload). Figure 4 shows the dimensionless pull-off force P^posubscript^𝑃𝑝𝑜\widehat{P}_{po}over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_p italic_o end_POSTSUBSCRIPT as a function of the dimensionless unloading rate r^=rτ/h0^𝑟𝑟𝜏subscript0\widehat{r}=r\tau/h_{0}over^ start_ARG italic_r end_ARG = italic_r italic_τ / italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for a material with n=0.8𝑛0.8n=0.8italic_n = 0.8 and by changing a^0=[1.94,2.50,3.61,5.28]subscript^𝑎01.942.503.615.28\widehat{a}_{0}=[1.94,2.50,3.61,5.28]over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = [ 1.94 , 2.50 , 3.61 , 5.28 ]. One realizes that while in the very beginning the enhancement curves all look similar, when the unloading rate starts to increase strong differences appear, with the curves referring to the smaller contact radius providing a much smaller enhancement, below the theoretical value even at the highest unloading rate tested.

Although, the dependence of the pull-off force on the initial contact radius and on the unloading rate has been discussed for the paradigmatic SLS model (Violano and Afferrante, 2022; Afferrante and Violano, 2022), it has remained unclear the importance of these parameters for real viscoelastic broad-band materials, hence we will address this question in the next subsection.

4.2 Threshold contact radius

Refer to caption
Refer to caption
Figure 5: Threshold contact radius: (a) Dimensionless pull-off force P^posubscript^𝑃𝑝𝑜\widehat{P}_{po}over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_p italic_o end_POSTSUBSCRIPT with respect to the normalized initial contact radius a^0subscript^𝑎0\widehat{a}_{0}over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT with the same unloading rate of r^=100.5^𝑟superscript100.5\widehat{r}=10^{0.5}over^ start_ARG italic_r end_ARG = 10 start_POSTSUPERSCRIPT 0.5 end_POSTSUPERSCRIPT for different material exponents n=[0.4,0.6,0.8,1.6]𝑛0.40.60.81.6n=[0.4,0.6,0.8,1.6]italic_n = [ 0.4 , 0.6 , 0.8 , 1.6 ], and k=0.1𝑘0.1k=0.1italic_k = 0.1; (b, inset) Dimensionless threshold contact radius a^0tsubscript^𝑎0𝑡\widehat{a}_{0t}over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT with respect to the normalized unloading rate r^^𝑟\widehat{r}over^ start_ARG italic_r end_ARG for different material exponents n=[0.4,0.6,0.8,1.6]𝑛0.40.60.81.6n=[0.4,0.6,0.8,1.6]italic_n = [ 0.4 , 0.6 , 0.8 , 1.6 ] and k=0.1𝑘0.1k=0.1italic_k = 0.1. (b, main figure) The same data reported in the inset are shown as dimensionless pull-off force P^posubscript^𝑃𝑝𝑜\widehat{P}_{po}over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_p italic_o end_POSTSUBSCRIPT with respect to the normalized threshold contact radius a^0tsubscript^𝑎0𝑡\widehat{a}_{0t}over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT. In all the panels triangle, star, diamond and square markers correspond respectively to n=[0.4,0.6,0.8,1.6]𝑛0.40.60.81.6n=[0.4,0.6,0.8,1.6]italic_n = [ 0.4 , 0.6 , 0.8 , 1.6 ].

Here, the results of a comprehensive numerical campaign specifically designed for providing, at a given unloading rate, the minimal initial contact radius (or preload) needed to maximize adhesion in a rigid sphere/soft substrate contact are provided. Figure 5 (a) shows the pull-off force obtained unloading the substrate at a given unloading rate r^=100.5^𝑟superscript100.5\widehat{r}=10^{0.5}over^ start_ARG italic_r end_ARG = 10 start_POSTSUPERSCRIPT 0.5 end_POSTSUPERSCRIPT as a function of the initial contact radius a^0subscript^𝑎0\widehat{a}_{0}over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for n=[0.4,0.6,0.8,1.6]𝑛0.40.60.81.6n=[0.4,0.6,0.8,1.6]italic_n = [ 0.4 , 0.6 , 0.8 , 1.6 ] (respectively triangle, star, diamond and square markers). The results show that as a^0subscript^𝑎0\widehat{a}_{0}over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT increases the pull-off force converges to a certain plateau and that in general, at a given r^^𝑟\widehat{r}over^ start_ARG italic_r end_ARG, broad-band materials (low n𝑛nitalic_n) will need a smaller initial contact radius to reach convergence of the pull-off force. Hence, in experiments, if the maximum adhesion is sought one must first perform a convergence study on the pre-loading conditions. In Fig. 5 (a) we have used a spline to interpolate the simulated points (markers), then we have computed the derivative dP^po/da^0𝑑subscript^𝑃𝑝𝑜𝑑subscript^𝑎0d\widehat{P}_{po}/d\widehat{a}_{0}italic_d over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_p italic_o end_POSTSUBSCRIPT / italic_d over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and set the condition dP^po/da^0<0.1𝑑subscript^𝑃𝑝𝑜𝑑subscript^𝑎00.1d\widehat{P}_{po}/d\widehat{a}_{0}<0.1italic_d over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_p italic_o end_POSTSUBSCRIPT / italic_d over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < 0.1 to determine a threshold contact radius indicated by a^0tsubscript^𝑎0𝑡\widehat{a}_{0t}over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT, above which we considered the pull-off force is converged. In Fig. 5 (a) the black curves change from dotted to solid when the contact radius is grater than a^0tsubscript^𝑎0𝑡\widehat{a}_{0t}over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT.

The results in Fig. 5 (a) refer to a particular unloading rate taken as a reference r^=100.5^𝑟superscript100.5\widehat{r}=10^{0.5}over^ start_ARG italic_r end_ARG = 10 start_POSTSUPERSCRIPT 0.5 end_POSTSUPERSCRIPT. A convergence study was performed over about 5555 orders of magnitude in terms of unloading rate, as shown in the inset of Fig. 5 (b), where every marker shown corresponds to the threshold contact radius a^0tsubscript^𝑎0𝑡\widehat{a}_{0t}over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT obtained for that material exponent n𝑛nitalic_n and at that given normalized unloading rate r^[100,106]^𝑟superscript100superscript106\widehat{r}\in[10^{0},10^{6}]over^ start_ARG italic_r end_ARG ∈ [ 10 start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ]. The inset of Fig. 5 (b) explicitly shows a dependence on the viscoelastic material spectrum broadness (i.e. the exponent n𝑛nitalic_n), nevertheless if the data are represented as normalized pull-off force at convergence as a function of a^0tsubscript^𝑎0𝑡\widehat{a}_{0t}over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT they collapse for all the exponents n𝑛nitalic_n into a single power law curve that we find to be P^po=0.3a^0t2.7subscript^𝑃𝑝𝑜0.3superscriptsubscript^𝑎0𝑡2.7\widehat{P}_{po}=0.3\widehat{a}_{0t}^{2.7}over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_p italic_o end_POSTSUBSCRIPT = 0.3 over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2.7 end_POSTSUPERSCRIPT (black dashed line in Fig. 5 (b)), which clearly saturates when the maximum enhancement P^po=P^JKR/k=1.5/ksubscript^𝑃𝑝𝑜subscript^𝑃𝐽𝐾𝑅𝑘1.5𝑘\widehat{P}_{po}=\widehat{P}_{JKR}/k=1.5/kover^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_p italic_o end_POSTSUBSCRIPT = over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_J italic_K italic_R end_POSTSUBSCRIPT / italic_k = 1.5 / italic_k is reached. Notice that, the smallest unloading rate considered in our analysis is r^3^𝑟3\widehat{r}\approx 3over^ start_ARG italic_r end_ARG ≈ 3 (see Fig. 5 (b), inset) as for quasi static unloading the elastic solution is retrieved and P^posubscript^𝑃𝑝𝑜\widehat{P}_{po}over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_p italic_o end_POSTSUBSCRIPT will not depend on a^0subscript^𝑎0\widehat{a}_{0}over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.

Furthermore Fig. 5 (b) shows that the transition from the power-law behaviour to the plateau is faster for materials with large material exponent n𝑛nitalic_n than for those characterized by low values of n𝑛nitalic_n, as a consequence of their narrow spectrum. Hence Fig. 5 (b) shows that regardless of the material model, the key parameter that determines the minimum contact radius a^0tsubscript^𝑎0𝑡\widehat{a}_{0t}over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT is the maximum amplification of the pull-off force that has to be reached.

5 Persson and Brener crack propagation theory for broad-band viscoelastic materials

In the previous sections, we have shown how the pull-off force depends on the unloading rate and on the initial contact area for various exponent n𝑛nitalic_n that characterize the broadness of the viscoelastic material response spectrum. Here, closed form solutions are obtained for the effective surface energy ΔγeffΔsubscript𝛾𝑒𝑓𝑓\Delta\gamma_{eff}roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT based on Persson and Brener (2005) crack propagation theory. It is useful to recall that for a Hertzian indenter, in the case of soft materials, the JKR model (Johnson et al., 1971) applies, which provides the pull-off force depends only on the sphere radius and surface energy PJKR=32πRΔγsubscript𝑃𝐽𝐾𝑅32𝜋𝑅Δ𝛾P_{JKR}=\frac{3}{2}\pi R\Delta\gammaitalic_P start_POSTSUBSCRIPT italic_J italic_K italic_R end_POSTSUBSCRIPT = divide start_ARG 3 end_ARG start_ARG 2 end_ARG italic_π italic_R roman_Δ italic_γ, hence, in the following, the normalized effective surface energy will be simply defined as Γ^eff=Δγeff/Δγ0Ppo/PJKRsubscript^Γ𝑒𝑓𝑓Δsubscript𝛾𝑒𝑓𝑓Δsubscript𝛾0similar-to-or-equalssubscript𝑃𝑝𝑜subscript𝑃𝐽𝐾𝑅\widehat{\Gamma}_{eff}=\Delta\gamma_{eff}/\Delta\gamma_{0}\simeq P_{po}/P_{JKR}over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT = roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT / roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≃ italic_P start_POSTSUBSCRIPT italic_p italic_o end_POSTSUBSCRIPT / italic_P start_POSTSUBSCRIPT italic_J italic_K italic_R end_POSTSUBSCRIPT.

We note that our crack propagation formulation is the extension of Persson and Brener (2005) idea of equating the input power from the remote load to the power that is dissipated due to the generation of new surfaces and due to viscoelastic dissipation, so one can obtain the effective surface energy ΔγeffΔsubscript𝛾𝑒𝑓𝑓\Delta\gamma_{eff}roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT as (Persson and Brener, 2005):

ΔγeffΔγ0Δsubscript𝛾𝑒𝑓𝑓Δsubscript𝛾0\displaystyle\frac{\Delta\gamma_{eff}}{\Delta\gamma_{0}}divide start_ARG roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT end_ARG start_ARG roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG =[1(1E0E)0+L(τ)(1/E01/E)τ{1+b2(τ)b1(τ)}𝑑τ]1,absentsuperscriptdelimited-[]11subscript𝐸0subscript𝐸superscriptsubscript0𝐿𝜏1subscript𝐸01subscript𝐸𝜏1superscript𝑏2𝜏superscript𝑏1𝜏differential-d𝜏1\displaystyle=\left[1-\left(1-\frac{E_{0}}{E_{\infty}}\right)\int_{0}^{+\infty% }\frac{L\left(\tau\right)}{\left(1/E_{0}-1/E_{\infty}\right)\tau}\left\{\sqrt{% 1+b^{-2}\left(\tau\right)}-b^{-1}\left(\tau\right)\right\}d\tau\right]^{-1}\;,= [ 1 - ( 1 - divide start_ARG italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG ) ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT divide start_ARG italic_L ( italic_τ ) end_ARG start_ARG ( 1 / italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 / italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ) italic_τ end_ARG { square-root start_ARG 1 + italic_b start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ( italic_τ ) end_ARG - italic_b start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_τ ) } italic_d italic_τ ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , (19)
b(τ)𝑏𝜏\displaystyle b\left(\tau\right)italic_b ( italic_τ ) =2πvτl0(Δγ0Δγeff).absent2𝜋𝑣𝜏subscript𝑙0Δsubscript𝛾0Δsubscript𝛾𝑒𝑓𝑓\displaystyle=\frac{2\pi v\tau}{l_{0}}\left(\frac{\Delta\gamma_{0}}{\Delta% \gamma_{eff}}\right)\;.= divide start_ARG 2 italic_π italic_v italic_τ end_ARG start_ARG italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ( divide start_ARG roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT end_ARG ) . (20)
Refer to caption
Figure 6: Normalized effective surface energy Γ^effsubscript^Γ𝑒𝑓𝑓\widehat{\Gamma}_{eff}over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT with respect to the normalized crack velocity v^^𝑣\widehat{v}over^ start_ARG italic_v end_ARG for different power law material exponent n=[0.4,0.6,0.8,1.6]𝑛0.40.60.81.6n=[0.4,0.6,0.8,1.6]italic_n = [ 0.4 , 0.6 , 0.8 , 1.6 ], respectively triangle, star, diamond and square markers, and k=0.1𝑘0.1k=0.1italic_k = 0.1. The blue circle markers in the plot correspond to the SLS material. Solid lines stand for the PB model (Eq. 23). The blue dashed line is a guide to the eye, showing the power law behaviour of the function Γ^eff(v^)=βv^msubscript^Γ𝑒𝑓𝑓^𝑣𝛽superscript^𝑣𝑚\widehat{\Gamma}_{eff}(\widehat{v})=\beta\widehat{v}^{m}over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ( over^ start_ARG italic_v end_ARG ) = italic_β over^ start_ARG italic_v end_ARG start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT in the intermediate velocity range. The inset depicts the fitting parameters {β,m}𝛽𝑚\{\beta,m\}{ italic_β , italic_m } for the values of n𝑛nitalic_n tested.

Introducing the dimensionless parameters:

v^=vτ0l0;τ^=ττ0,formulae-sequence^𝑣𝑣subscript𝜏0subscript𝑙0^𝜏𝜏subscript𝜏0\widehat{v}=\frac{v\tau_{0}}{l_{0}};\qquad\widehat{\tau}=\frac{\tau}{\tau_{0}}\;,over^ start_ARG italic_v end_ARG = divide start_ARG italic_v italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ; over^ start_ARG italic_τ end_ARG = divide start_ARG italic_τ end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG , (21)

and substituting the retardation spectrum defined for the MPL material model in Eq. (8) into Eq. (19) one gets

Γ^eff=[1(1k)0+τ^n1exp(τ^)Γ(n)[1+(Γ^eff2πv^1τ^)2(Γ^eff2πv^1τ^)]𝑑τ^]1,subscript^Γ𝑒𝑓𝑓superscriptdelimited-[]11𝑘superscriptsubscript0superscript^𝜏𝑛1^𝜏Γ𝑛delimited-[]1superscriptsubscript^Γ𝑒𝑓𝑓2𝜋^𝑣1^𝜏2subscript^Γ𝑒𝑓𝑓2𝜋^𝑣1^𝜏differential-d^𝜏1\widehat{\Gamma}_{eff}=\left[1-\left(1-k\right)\int_{0}^{+\infty}\frac{% \widehat{\tau}^{n-1}\exp\left(-\widehat{\tau}\right)}{\Gamma\left(n\right)}% \left[\sqrt{1+\left(\frac{\widehat{\Gamma}_{eff}}{2\pi\widehat{v}}\frac{1}{% \widehat{\tau}}\right)^{2}}-\left(\frac{\widehat{\Gamma}_{eff}}{2\pi\widehat{v% }}\frac{1}{\widehat{\tau}}\right)\right]d\widehat{\tau}\right]^{-1}\;,over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT = [ 1 - ( 1 - italic_k ) ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT divide start_ARG over^ start_ARG italic_τ end_ARG start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT roman_exp ( - over^ start_ARG italic_τ end_ARG ) end_ARG start_ARG roman_Γ ( italic_n ) end_ARG [ square-root start_ARG 1 + ( divide start_ARG over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_π over^ start_ARG italic_v end_ARG end_ARG divide start_ARG 1 end_ARG start_ARG over^ start_ARG italic_τ end_ARG end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - ( divide start_ARG over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_π over^ start_ARG italic_v end_ARG end_ARG divide start_ARG 1 end_ARG start_ARG over^ start_ARG italic_τ end_ARG end_ARG ) ] italic_d over^ start_ARG italic_τ end_ARG ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , (22)

which can be written as

Γ^eff=[1(1k)I(n,v^,Γ^eff)]1,subscript^Γ𝑒𝑓𝑓superscriptdelimited-[]11𝑘𝐼𝑛^𝑣subscript^Γ𝑒𝑓𝑓1\widehat{\Gamma}_{eff}=\left[1-\left(1-k\right)I\left(n,\widehat{v},\widehat{% \Gamma}_{eff}\right)\right]^{-1}\,,over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT = [ 1 - ( 1 - italic_k ) italic_I ( italic_n , over^ start_ARG italic_v end_ARG , over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , (23)

where I(n,v^,Γ^eff)𝐼𝑛^𝑣subscript^Γ𝑒𝑓𝑓I(n,\widehat{v},\widehat{\Gamma}_{eff})italic_I ( italic_n , over^ start_ARG italic_v end_ARG , over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) stands for the integral in Eq. 22 which expression is given in closed form in the B. Following Persson and Brener (2005) original arguments, we determine the lengthscale l0subscript𝑙0l_{0}italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT equating the linear elastic fracture mechanics stress field to the critical stress σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT required to break the atomic bonds. Hence:

σcsubscript𝜎𝑐\displaystyle\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT =KI2πl0;KI2=Δγ02E0,formulae-sequenceabsentsubscript𝐾𝐼2𝜋subscript𝑙0superscriptsubscript𝐾𝐼2Δsubscript𝛾02superscriptsubscript𝐸0\displaystyle=\frac{K_{I}}{\sqrt{2\pi l_{0}}};\qquad K_{I}^{2}=\frac{\Delta% \gamma_{0}}{2E_{0}^{*}}\;,= divide start_ARG italic_K start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG 2 italic_π italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG end_ARG ; italic_K start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG , (24)
l0subscript𝑙0\displaystyle l_{0}italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT =E0Δγ0πσc2=E0Δγ0π(ασ0)2,absentsuperscriptsubscript𝐸0Δsubscript𝛾0𝜋superscriptsubscript𝜎𝑐2superscriptsubscript𝐸0Δsubscript𝛾0𝜋superscript𝛼subscript𝜎02\displaystyle=\frac{E_{0}^{{}^{*}}\Delta\gamma_{0}}{\pi\sigma_{c}^{2}}=\frac{E% _{0}^{{}^{*}}\Delta\gamma_{0}}{\pi\left(\alpha\sigma_{0}\right)^{2}}\;,= divide start_ARG italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ∗ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_π italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = divide start_ARG italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ∗ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_π ( italic_α italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (25)

where E0=E01ν2superscriptsubscript𝐸0subscript𝐸01superscript𝜈2E_{0}^{{}^{*}}=\frac{E_{0}}{1-\nu^{2}}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ∗ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT = divide start_ARG italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 1 - italic_ν start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG is the rubbery plain strain elastic modulus of the halfspace, KIsubscript𝐾𝐼K_{I}italic_K start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT is the stress intensity factor in mode I and the “2222” in its expression takes into account that one of the contacting bodies is rigid, while α𝛼\alphaitalic_α in Eq. (25) is a coefficient of order unity to relate the critical stress σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in PB theory to the σ0subscript𝜎0\sigma_{0}italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT we are using in the numerical simulations that are based on the LJ force-separation law. Notice that for soft polymers l0/h01subscript𝑙0subscript01{l_{0}}/{h_{0}}\approx 1italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 1 hence l0subscript𝑙0l_{0}italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT should physically be of the same order of the intermolecular distance.

Solving Eq. (23) for n=[0.2,0.4,0.6,0.8,1.6]𝑛0.20.40.60.81.6n=[0.2,0.4,0.6,0.8,1.6]italic_n = [ 0.2 , 0.4 , 0.6 , 0.8 , 1.6 ], k=0.1𝑘0.1k=0.1italic_k = 0.1 and for varying crack velocity v^^𝑣\widehat{v}over^ start_ARG italic_v end_ARG one easily find the results shown in Fig. 6 (black solid lines). So, for a given effective energy, broad-band materials would require a much higher crack speed than for narrow-band materials. The numerical results from the same set of parameters are shown in Fig. 6 as markers (n=[0.4,0.6,0.8,1.6]𝑛0.40.60.81.6n=[0.4,0.6,0.8,1.6]italic_n = [ 0.4 , 0.6 , 0.8 , 1.6 ], respectively triangle, star, diamond and square markers), where we find an excellent agreement with the analytical results by using α=π/90.3491𝛼𝜋9similar-to-or-equals0.3491\alpha=\pi/9\simeq 0.3491italic_α = italic_π / 9 ≃ 0.3491. It is reminded that the numerical results shown in Fig. 6 have been obtained unloading a fully relaxed halfspace and are related to an initial contact radius exceeding the threshold value, i.e a^0>a^0tsubscript^𝑎0subscript^𝑎0𝑡\widehat{a}_{0}>\widehat{a}_{0t}over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT (see Fig. 5). Numerical simulations conducted for k=E0/E=[0.01,0.05,0.1]𝑘subscript𝐸0subscript𝐸0.010.050.1k=E_{0}/E_{\infty}=[0.01,0.05,0.1]italic_k = italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = [ 0.01 , 0.05 , 0.1 ] confirmed that απ/9similar-to-or-equals𝛼𝜋9\alpha\simeq\pi/9italic_α ≃ italic_π / 9 independently on the ratio rubbery to glassy modulus k𝑘kitalic_k.

Refer to caption
Figure 7: (main figure) Normalized effective surface energy Γ^effsubscript^Γ𝑒𝑓𝑓\widehat{\Gamma}_{eff}over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT based on the numerical BEM simulations versus the normalized effective surface energy predicted by using the Eq.s (23,26). (inset) Normalized crack velocity v^^𝑣\widehat{v}over^ start_ARG italic_v end_ARG versus the normalized unloading rate (r^PBsubscript^𝑟𝑃𝐵\widehat{r}_{PB}over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_P italic_B end_POSTSUBSCRIPT). In both panels the same simulations results are shown, in particular for different power law material exponent n=[0.4,0.6,0.8,1.6]𝑛0.40.60.81.6n=[0.4,0.6,0.8,1.6]italic_n = [ 0.4 , 0.6 , 0.8 , 1.6 ], respectively triangle, star, diamond and square markers, and k=[0.01,0.05,0.1]𝑘0.010.050.1k=[0.01,0.05,0.1]italic_k = [ 0.01 , 0.05 , 0.1 ] respectively markers with a blue contour line, with a black contour line and without contour line. Blue circles stand for the SLS with k=0.1𝑘0.1k=0.1italic_k = 0.1.

It is worth mentioning that the SLS is very often used as a paradigmatic model for a polymer viscoelastic behavior. As a comparison, Fig. 6 reports the results obtained for a SLS as blue circles, which confirms the case of a SLS is close to n=1.6𝑛1.6n=1.6italic_n = 1.6 and shows a notably large amplification of interfacial adhesion at relatively low crack speed if it is compared with broad-spectrum viscoelastic material. Our experimental results will show in Section 6 that 10:1 PDMS silicone have an exponent n0.22similar-to-or-equals𝑛0.22n\simeq 0.22italic_n ≃ 0.22, which implies the maximum adhesion amplification may be observed only at unloading rates which are orders of magnitude larger than that needed for a SLS, which poses also questions about the practical feasibility of reaching so large retraction rates and possible nonlinear effects that may come into play, which will be discussed in the Discussion section. For a more convenient use of Eq. (23), the power law scaling of the effective surface energy in the intermediate velocity range is reported here as Γ^eff=βv^msubscript^Γ𝑒𝑓𝑓𝛽superscript^𝑣𝑚\widehat{\Gamma}_{eff}=\beta\widehat{v}^{m}over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT = italic_β over^ start_ARG italic_v end_ARG start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT (see blue dashed line in Fig. 6), where the parameters β,m𝛽𝑚{\beta,m}italic_β , italic_m can be found in Fig. 6 inset.

The applicability of Eq. (23) for the prediction of the effective surface energy would remain limited by the fact that in all the viscoelastic crack propagation theories, including Eq. (23), the enhancement of the surface energy is a function of the crack velocity at pull-off which is generally not an input parameter in experiments and would be anyway difficult to control. Nevertheless, Fig. 7 shows in the inset that the crack velocity at pull-off v^^𝑣\widehat{v}over^ start_ARG italic_v end_ARG scales approximately as

v^=2.887r^PB1.171,^𝑣2.887superscriptsubscript^𝑟𝑃𝐵1.171\widehat{v}=2.887\widehat{r}_{PB}^{1.171}\;,over^ start_ARG italic_v end_ARG = 2.887 over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_P italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1.171 end_POSTSUPERSCRIPT , (26)

over about 10 orders of magnitude in term of unloading rate r^PBsubscript^𝑟𝑃𝐵\widehat{r}_{PB}over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_P italic_B end_POSTSUBSCRIPT, where r^PB=rτ/l0subscript^𝑟𝑃𝐵𝑟𝜏subscript𝑙0\widehat{r}_{PB}=r\tau/l_{0}over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_P italic_B end_POSTSUBSCRIPT = italic_r italic_τ / italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Figure 7 shows the numerical results obtained for the material exponents n=[0.4,0.6,0.8,1.6]𝑛0.40.60.81.6n=[0.4,0.6,0.8,1.6]italic_n = [ 0.4 , 0.6 , 0.8 , 1.6 ], respectively triangles, stars, diamonds, squares (circles stand for the SLS material) and for k=[0.01,0.05,0.1]𝑘0.010.050.1k=[0.01,0.05,0.1]italic_k = [ 0.01 , 0.05 , 0.1 ] respectively markers with a blue contour line, with a black contour line and without contour line. Filled blue circles stand for the SLS with k=0.1𝑘0.1k=0.1italic_k = 0.1. Hence by using the Eq. (26) to estimate the crack speed at pull-off as a function of the retraction rate we have used Eq. (23) to predict the effective surface energy and compared with the numerical BEM results, which using the same symbols as in the inset, are shown in the main Fig. 7. The solid black line represents the condition of perfect match between prediction and actual numerical results, while as a guide to the eye we have drawn also two dashed lines representing ±15%plus-or-minuspercent15\pm 15\%± 15 % error. Although the scaling may be improved by using more refined models, the use of Eq.s (23,26) makes the estimate of the pull-off force straightforward based only on the material parameters and on the unloading rate. It is recalled that all the numerical results have been obtained for the Tabor parameter μ=3.24𝜇3.24\mu=3.24italic_μ = 3.24, hence we expect Eq. (26) to be valid in the limit of short-range adhesion also referred to as the ”JKR limit” (Johnson et al., 1971).

6 Experimental adhesion tests

In the previous sections we have developed a general MPL material model capable of describing the viscoelastic behaviour of both narrow and broad band materials, then we have compared BEM numerical results with PB theory finding an excellent agreement. Finally, in this section the numerical predictions will be validated against experimental results.

A series of adhesion tests where performed using a smooth spherical lens loaded and unloaded from a soft viscoelastic substrate at various unloading velocities. The spherical lens was made of borosilicate crown glass (SLB-05-10P, Sigma Koki) with a nominal radius of R = 5.19 mm and the substrates were made of polydimethylsiloxane (PDMS, Sylgard 184, DowCorning Corporation) with resin to curing agent weight ratio of 10:1. PDMS is a silicone elastomer well known to exhibit viscoelastic properties, as confirmed in several previous studies (Lorenz et al., 2013; VanDonselaar et al., 2023; Violano et al., 2021; Petroli et al., 2022). For the substrate material characterization a classical dog-bone shaped specimen was fabricated and used for dynamic mechanical analysis (DMA). All the samples were cured at 70 C for two hours on a heating table and then followed by natural cooling.

6.1 Material characterization

Refer to caption
Figure 8: (a) Real part of the complex elastic modulus Esuperscript𝐸E^{\prime}italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT in the frequency domain at Tamb=20T_{amb}=20\;{{}^{\circ}}italic_T start_POSTSUBSCRIPT italic_a italic_m italic_b end_POSTSUBSCRIPT = 20 start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPTC. (b) Imaginary part E′′superscript𝐸′′E^{\prime\prime}italic_E start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT of the complex elastic modulus in the frequency domain at Tamb=20subscript𝑇𝑎𝑚𝑏20T_{amb}=20italic_T start_POSTSUBSCRIPT italic_a italic_m italic_b end_POSTSUBSCRIPT = 20 C. In both panels: the black curve with circle markers stands for the experimental data, the blue curve for the fitted MPL material model and the red curve for the fitted GMM model with 18 arms (see Fig. 11).

The DMA test was performed using a DMA850 (TA Instruments) to characterize the viscoelastic properties of the PDMS. The dog-bone-shaped specimen had cross-sectional dimensions of 3.863.863.863.86 mm in width and 0.750.750.750.75 mm in thickness. Temperature sweeps were conducted at a fixed frequency of f=1𝑓1f=1italic_f = 1 Hz and a strain amplitude of ϵ=0.1%italic-ϵpercent0.1\epsilon=0.1\%italic_ϵ = 0.1 %. The temperature runs from 130-130\;{{}^{\circ}}- 130 start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPTC to 2020\;{{}^{\circ}}20 start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPTC with 1010\;{{}^{\circ}}10 start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPTC step size. To move from temperature to frequency domain we used the WLF time-temperature superposition (Williams et al., 1955), hence the shift factor is defined as

log10aT=log10fTgfT=17.44(TTg)51.6+TTg,subscript10subscript𝑎𝑇subscript10subscript𝑓subscript𝑇𝑔subscript𝑓𝑇17.44𝑇subscript𝑇𝑔51.6𝑇subscript𝑇𝑔\log_{10}a_{T}=\log_{10}\frac{f_{T_{g}}}{f_{T}}=\frac{-17.44\left(T-T_{g}% \right)}{51.6+T-T_{g}}\,,roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT divide start_ARG italic_f start_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG italic_f start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_ARG = divide start_ARG - 17.44 ( italic_T - italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ) end_ARG start_ARG 51.6 + italic_T - italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG , (27)

where fTsubscript𝑓𝑇f_{T}italic_f start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT is the frequency at the temperature T𝑇Titalic_T and Tgsubscript𝑇𝑔T_{g}italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT is the glass transition temperature. For the PDMS substrate we assumed Tg=115subscript𝑇𝑔superscript115T_{g}=-115^{\circ}italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = - 115 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPTC, which agrees well with the results reported in Ref. (VanDonselaar et al., 2023) for the same material. Furthermore, we note that using Tg=115subscript𝑇𝑔superscript115T_{g}=-115^{\circ}italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = - 115 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPTC our measurements of the complex modulus E¯¯𝐸\overline{E}over¯ start_ARG italic_E end_ARG also satisfy the Kramers-Kronig (KK) relation (Pritz, 2005)

E′′(ω)=2ωπ0+E(u)ω2u2𝑑u,superscript𝐸′′𝜔2𝜔𝜋superscriptsubscript0superscript𝐸𝑢superscript𝜔2superscript𝑢2differential-d𝑢E^{\prime\prime}\left(\omega\right)=-\frac{2\omega}{\pi}\int_{0}^{+\infty}% \frac{E^{\prime}\left(u\right)}{\omega^{2}-u^{2}}du\,,italic_E start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ω ) = - divide start_ARG 2 italic_ω end_ARG start_ARG italic_π end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT divide start_ARG italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_u ) end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_d italic_u , (28)

where ω=2πf𝜔2𝜋𝑓\omega=2\pi fitalic_ω = 2 italic_π italic_f is the angular frequency and the integral should be intended as its Principal Value (Pritz, 2005).

The experimental data for the complex modulus were shifted to Tamb=20subscript𝑇𝑎𝑚𝑏superscript20T_{amb}=20\;^{\circ}italic_T start_POSTSUBSCRIPT italic_a italic_m italic_b end_POSTSUBSCRIPT = 20 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPTC by using Eq. (27) and fitted using Eq. (9), which is written in terms of the complex compliance C(ω)𝐶𝜔C(\omega)italic_C ( italic_ω ) as that is the function needed in the numerical BEM implementation (see Eq. (17)). Figure 8, panels (a)-(b), shows the complex modulus E¯(ω)=1/C¯(ω)=E(ω)+𝒊E′′(ω)¯𝐸𝜔1¯𝐶𝜔superscript𝐸𝜔𝒊superscript𝐸′′𝜔\overline{E}\left(\omega\right)=1/\overline{C}\left(\omega\right)=E^{\prime}% \left(\omega\right)+\boldsymbol{i}E^{\prime\prime}\left(\omega\right)over¯ start_ARG italic_E end_ARG ( italic_ω ) = 1 / over¯ start_ARG italic_C end_ARG ( italic_ω ) = italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω ) + bold_italic_i italic_E start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ω ) as obtained experimentally (black solid curve with circle markers) and as fitted by the MPL material model (blue solid curve). For PDMS we found

E0=1.458 MPaE=3.089103 MPan=0.2207τ0=0.01876 ssubscript𝐸01.458 MPasubscript𝐸3.089superscript103 MPa𝑛0.2207subscript𝜏00.01876 s\begin{array}[c]{c}E_{0}=1.458\text{ MPa}\\ E_{\infty}=3.089\ast 10^{3}\text{ MPa}\\ n=0.2207\\ \tau_{0}=0.01876\text{ s}\end{array}start_ARRAY start_ROW start_CELL italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1.458 MPa end_CELL end_ROW start_ROW start_CELL italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 3.089 ∗ 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT MPa end_CELL end_ROW start_ROW start_CELL italic_n = 0.2207 end_CELL end_ROW start_ROW start_CELL italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.01876 s end_CELL end_ROW end_ARRAY (29)

For comparison purpose, the result that would have been obtained by fitting the experimental data using a Generalized Maxwell Model (GMM, also known as the ”Wiechert model”) with 18 arms, hence 37 constants, is also shown in Fig. 8 as a dashed red curve. One realizes that both the GMM and the MPL models give a fair representation of the material behavior, although the MPL model is simpler to use, and the four parameters used in the fitting procedure {E0,E,n,τ0}subscript𝐸0subscript𝐸𝑛subscript𝜏0\left\{E_{0},E_{\infty},n,\tau_{0}\right\}{ italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT , italic_n , italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT } have a straightforward physical interpretation.

6.2 Experimental setup and comparison

A custom-designed adhesion test instrument, based on the tribometer platform (NTR2, CSM Instruments), was constructed to measure the pull-off force. As illustrated in Fig. 9 (a), the lens was rigidly fixed to the force sensor. The PDMS substrate was positioned above a transparent rigid box, with the contact interface observable through a camera via a prism mounted inside the box. The pull-off tests comprised three sequential steps: loading, dwelling, and unloading. Initially, the lens was gradually loaded against the PDMS substrate with a preload force denoted as P𝑃Pitalic_P, followed by a dwell period of 60 seconds to ensure complete relaxation of adhesive contact. Subsequently, the lens was pulled out at a fixed unloading rate, r𝑟ritalic_r. Throughout the entire process, the normal force was recorded and the pull-off force represents the absolute minimum normal force. Firstly, we measured the variation of the normal force with the contact radius, a𝑎aitalic_a, at a very low unloading rate r=0.98μ𝑟0.98𝜇r=0.98\muitalic_r = 0.98 italic_μm/s to determine the interfacial parameters, as shown in Fig. 9 (b). By fitting the relationship between the normal force and contact radius using Carpick’s method (Carpick et al., 1999), we estimated the intrinsic work of adhesion Δγ=0.152J/m2Δ𝛾0.152superscriptJ/m2{\Delta\gamma}=0.152\;\text{J/m}^{2}roman_Δ italic_γ = 0.152 J/m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and the Tabor parameter μ=2.05𝜇2.05\mu=2.05italic_μ = 2.05. Next, we conducted tests by varying the unloading rate r𝑟ritalic_r. The lens is brought into contact with the PDMS substrate and loaded to the preset preload P0=1.5subscript𝑃01.5P_{0}=1.5italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1.5 mN. After a 60-second dwell period, the lens is moved upward until the contact is broken and the lens is pulled off from the substrate.

Refer to caption
(a)
Refer to caption
(b)
Figure 9: (a) Schematic of experimental setup for adhesion tests; (b) Variations of normal force with contact radius a𝑎aitalic_a at a very low unloading rate to determine the interfacial parameters.

We used our numerical BEM code, using the MPL material model for the viscoelastic substrate, to predict the pull-off force during the unloading process. The comparison with experiments leads to the result shown in Fig. 10, where the pull-off force, Pposubscript𝑃𝑝𝑜{P}_{po}italic_P start_POSTSUBSCRIPT italic_p italic_o end_POSTSUBSCRIPT [mN], is plotted as a function of the unloading rate, r𝑟ritalic_r [μ𝜇\muitalic_μm/s] (the red squares stand for the experimental data, the black solid line for the numerical results). According to Fig. 10, the experimental results confirm an increase in the pull-off force with increasing unloading rates. While there is good agreement between numerical and experimental results in the range of retraction rates r=[1,100]𝑟1100r=[1,100]italic_r = [ 1 , 100 ] μ𝜇\muitalic_μm/s, the experimental data exhibit marked higher values than the numerical predictions for high values of the unloading rate r>100𝑟100r>100italic_r > 100 μ𝜇\muitalic_μm/s, which agrees well with other published experimental results (VanDonselaar et al., 2023; Tiwari et al., 2017). For having a good fit of the low speed experimental results, we set h0=30.8subscript030.8h_{0}=30.8italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 30.8 nm, which is discussed in detail in the Discussion section.

Refer to caption
Refer to caption
Figure 10: (a) Pull-off force as a function of the unloading rate: comparison between numerical (from the BEM code, solid black line, labelled (Ppo)linsubscriptsubscript𝑃𝑝𝑜𝑙𝑖𝑛(P_{po})_{lin}( italic_P start_POSTSUBSCRIPT italic_p italic_o end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_l italic_i italic_n end_POSTSUBSCRIPT) and experimental (red square markers) results for preload P0=1.5subscript𝑃01.5P_{0}=1.5italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1.5 mN, R=5.19𝑅5.19R=5.19italic_R = 5.19 m-3, Δγ0=152.3Δsubscript𝛾0152.3\Delta\gamma_{0}=152.3roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 152.3 mJ/m2, h0=30.8subscript030.8h_{0}=30.8italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 30.8 nm. The green dashed line was obtained using the effective surface energy fitted on the experimental results labelled as (Δγeff)GSsubscriptΔsubscript𝛾𝑒𝑓𝑓𝐺𝑆(\Delta\gamma_{eff})_{GS}( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_G italic_S end_POSTSUBSCRIPT and shown in panel (b). (b, left y-axis) Effective surface energy from: the experimental results (red squares, labelled (Δγeff)expsubscriptΔsubscript𝛾𝑒𝑓𝑓𝑒𝑥𝑝(\Delta\gamma_{eff})_{exp}( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_e italic_x italic_p end_POSTSUBSCRIPT), the fit of the experimental data using a Gent and Schulz power law model (Eq. (30), where v0=213.5subscript𝑣0213.5v_{0}=213.5italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 213.5 μ𝜇\muitalic_μm/s and ξ=0.4154𝜉0.4154\xi=0.4154italic_ξ = 0.4154, dashed green line, labelled (Δγeff)GSsubscriptΔsubscript𝛾𝑒𝑓𝑓𝐺𝑆(\Delta\gamma_{eff})_{GS}( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_G italic_S end_POSTSUBSCRIPT), the prediction obtained for the PDMS substrate using the linear numerical BEM model (solid black line, labelled (Δγeff)linsubscriptΔsubscript𝛾𝑒𝑓𝑓𝑙𝑖𝑛(\Delta\gamma_{eff})_{lin}( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_l italic_i italic_n end_POSTSUBSCRIPT). In this respect we used our approximate Eq. (26) to determine the crack velocity at pull-off starting from the experimental retraction rate. (b, right y-axis) The ratio between (Δγeff)GS/(Δγeff)linsubscriptΔsubscript𝛾𝑒𝑓𝑓𝐺𝑆subscriptΔsubscript𝛾𝑒𝑓𝑓𝑙𝑖𝑛(\Delta\gamma_{eff})_{GS}/(\Delta\gamma_{eff})_{lin}( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_G italic_S end_POSTSUBSCRIPT / ( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_l italic_i italic_n end_POSTSUBSCRIPT for which a power law fit is provided (Δγeff)GS(Δγeff)lin=1+(vv1)1/2subscriptΔsubscript𝛾𝑒𝑓𝑓𝐺𝑆subscriptΔsubscript𝛾𝑒𝑓𝑓𝑙𝑖𝑛1superscript𝑣subscript𝑣112\frac{(\Delta\gamma_{eff})_{GS}}{(\Delta\gamma_{eff})_{lin}}=1+(\frac{v}{v_{1}% })^{1/2}divide start_ARG ( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_G italic_S end_POSTSUBSCRIPT end_ARG start_ARG ( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_l italic_i italic_n end_POSTSUBSCRIPT end_ARG = 1 + ( divide start_ARG italic_v end_ARG start_ARG italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT where v1=104μsubscript𝑣1superscript104𝜇v_{1}=10^{4}\muitalic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_μm/s (dashed pale blue curve).

7 Discussion

7.1 On the possibility to reach the maximum adhesion enhancement

According to viscoelastic crack propagation theories (Persson and Brener, 2005; Schapery, 1975a, b) the maximum enhancement of the pull-off force is P^po=P^JKR/ksubscript^𝑃𝑝𝑜subscript^𝑃𝐽𝐾𝑅𝑘\widehat{P}_{po}=\widehat{P}_{JKR}/kover^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_p italic_o end_POSTSUBSCRIPT = over^ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_J italic_K italic_R end_POSTSUBSCRIPT / italic_k, hence based on the results reported in Section 4, one can estimate that for a PDMS material with k4.73104similar-to-or-equals𝑘4.73superscript104k\simeq 4.73*10^{-4}italic_k ≃ 4.73 ∗ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT the maximum amplification of the pull-off force will be reached for a^0t30.9similar-to-or-equalssubscript^𝑎0𝑡30.9\widehat{a}_{0t}\simeq 30.9over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT ≃ 30.9. Using the interfacial and material properties we have found for PDMS (Δγ0=152.3 mJ/m2,ν=0.5,E01.94 MPa)formulae-sequenceΔsubscript𝛾0152.3superscript mJ/m2formulae-sequence𝜈0.5similar-to-or-equalssubscriptsuperscript𝐸01.94 MPa(\Delta\gamma_{0}=152.3\textrm{ mJ/m}^{2},\nu=0.5,E^{*}_{0}\simeq 1.94\textrm{% MPa})( roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 152.3 mJ/m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , italic_ν = 0.5 , italic_E start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≃ 1.94 MPa ) and for R=5.19 mm𝑅5.19 mmR=5.19\textrm{ mm}italic_R = 5.19 mm gives an initial contact radius of a0t=5.8subscript𝑎0𝑡5.8{a}_{0t}=5.8italic_a start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT = 5.8 mm, which is larger then the sphere radius and even considering a parabolic (Hertzian) profile certainly outside the limit of validity of the hypothesis of small deformations, which raises doubts about the practical feasibility of reaching the maximum amplification factor predicted by crack propagation theories.

Another consideration to be made is related to the unloading rate that would be needed to reach the maximum adhesion amplification. By using the results reported in Section 6, one estimates that to reach the maximum amplification for a PDMS substrate one would need to unload the substrate at v^109^𝑣superscript109\widehat{v}\approx 10^{9}over^ start_ARG italic_v end_ARG ≈ 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT (see Fig. 6), which using {l0=30.0 nm,τ0=0.01876 s}formulae-sequencesubscript𝑙030.0 nmsubscript𝜏00.01876 s\{l_{0}=30.0\textrm{ nm},\tau_{0}=0.01876\textrm{ s}\}{ italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 30.0 nm , italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.01876 s } together with Eq. (26) gives the dimensional unloading rate of about r31.4𝑟31.4r\approx 31.4italic_r ≈ 31.4 m/s, which is about 4 orders of magnitude larger than the maximum unloading velocity usually used in adhesion experiments (Tiwari et al., 2017; VanDonselaar et al., 2023), provided also the limitations introduced by the inertia of the motorized linear stages. Hence, insufficient preload and unloading rates used in experiments may partially explain why, in Literature, measurements of very large enhancement factors, even close to 1/k1𝑘1/k1 / italic_k, are missing (see for example VanDonselaar et al. (2023); Tiwari et al. (2017)).

On the other hand, our numerical and experimental results seem to be in agreement with the experimental adhesion tests reported in Refs. (VanDonselaar et al., 2023; Tiwari et al., 2017) for a similar PDMS material, where they also found that PB theory agreed well with experimental observations only up to about r100 μm/s𝑟100 μ𝑚𝑠r\approx 100\textrm{ $\mu$}m/sitalic_r ≈ 100 italic_μ italic_m / italic_s. This may suggest that both numerical and theoretical models are lacking of essential phenomena to describe the detachment process at high retraction rates. At present, a few hypotheses have been formulated, ranging from the possibility of nonlinear dissipative phenomena, happening within the process zone, like (i) cavitation and stringing, (ii) extraction of non-cross linked polymeric chains from the substrate, (iii) temperature dependence of the material behaviour at the crack tip, (iv) the nonlinear behaviour of the material at the large strains (10%absentpercent10\approx 10\%≈ 10 %) experienced close to the crack tip VanDonselaar et al. (2023); Tiwari et al. (2017), of course not included into the (linear) theoretical and numerical models, which is discussed in the next subsection.

7.2 Energy dissipation within the process zone

To obtain a satisfactory fit of the experimental data at low unloading rates we set h0=30.8subscript030.8h_{0}=30.8italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 30.8 nm. Notice that from quasi-static experiments, using the definition of the Tabor parameter we would have obtained a much larger equilibrium distance h0=1.55subscript01.55h_{0}=1.55italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1.55 μm𝜇m\mu\textrm{m}italic_μ m, which is close to what can be obtained for the same PDMS material using the parameters in Oliver et al. (2023). For the PDMS material we have characterized, using h0=30.8subscript030.8h_{0}=30.8italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 30.8 nm, one obtains that the size of the process zone that fits the experimental data in Fig. 10a is l030.0similar-to-or-equalssubscript𝑙030.0l_{0}\simeq 30.0italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≃ 30.0 nm.

Indeed, determining the length of the process zone in viscoelastic crack propagation is still an open question. In the de Gennes (1996) and Saulnier et al. (2004) theories, the size of the ”nonlinear” zone is assumed to be a constant, and the fracture energy has its maximum amplification at intermediate speeds. In PB theory this size is not constant and is directly proportional to the applied energy release rate G𝐺Gitalic_G, which results in a model practically coincident with the cohesive zone model of Knauss and Schapery (see Knauss (2015)). However, in fitting experimental data of fracture Hui et al. (2022) consider two examples, a styrene-butadiene co-polymer from Gent and Lai (1994), where they don’t have independent estimate of the cohesive strength, but simply fit the fracture energy vs speed data, obtaining a process zone size at low speed of a nonphysical size of 0.10.10.10.1 nm, consistent with Gent and Lai (1994). In the second example, they consider a polyurethane elastomer called Solithane 113 of Knauss (2015), and obtain by the same process a size of 1111 nm. Hence, Hui et al. (2022) conclude that this size cannot realistically represent a dissipation zone for which a lower bound should be the length of the monomer unit 46absent46\approx 46≈ 46 nm (Lake and Thomas, 1967). 222Recent literature contributions have started to question the validity of classical linear elastic fracture mechanics for unfilled plastics and elastomers suggesting that fracture initiates at a critical tensile strength, see Wang et al. (2024). Notice that non linear crack propagation theories have been developed by Schapery using cohesive models (Schapery, 1984), and have provided a fracture process zone at low speeds of approximately 10101010 nm, much more realistic than the 0.10.10.10.1 nm found by Knauss (2015) and Schapery (1975b) for Solithane rubber. In fitting crack propagation data in rubbers, Schapery (2022a) (Tab. 1) found a jump in propagation speed at a certain applied load which seems to suggest a sharp change of cohesive zone fracture energy as function of speed. He found a low speed fracture energy which is higher than the fast propagation speed fracture energy of a factor of about 6. In our adhesion experiments, our theory is linear and hence we cannot exclude that a non-linear theory would explain this apparent continuous change of cohesive zone fracture energy with speed of the linear theory, which is an increase with speed rather than a decrease and hence gives no instability.

Barthel (2024) reports post mortem experimental measurements of the process zone length from damage occurred at the crack tip and shows this should be of a physically reasonable size of the order of microns. Also, it clearly increases with size as PB and Schapery suggest, but contrary to the original DeGennes and Sauliner theories. Furthermore, linear theories seems to work better for very viscoelastic solids, namely when the glass transition temperature is above ambient temperature, perhaps because for very viscoelastic materials the dissipation in the bulk becomes dominant (Barthel, 2024).

We have estimated the experimental effective surface energy Δγeff(v)Δsubscript𝛾𝑒𝑓𝑓𝑣\Delta\gamma_{eff}(v)roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ( italic_v ) (Fig. 10b, left y-axis) as obtained from experiments (red squares), fitted by a Gent-Schultz (Gent and Schultz, 1972) power law equation (green dashed curve)

Δγeff=Δγ0(1+(vv0)ξ),Δsubscript𝛾𝑒𝑓𝑓Δsubscript𝛾01superscript𝑣subscript𝑣0𝜉\Delta\gamma_{eff}=\Delta\gamma_{0}\left(1+\left(\frac{v}{v_{0}}\right)^{\xi}% \right)\;,roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT = roman_Δ italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 + ( divide start_ARG italic_v end_ARG start_ARG italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_ξ end_POSTSUPERSCRIPT ) , (30)

and estimated from our linear BEM numerical scheme (black solid line), respectively {(Δγeff)exp,(Δγeff)GS,(Δγeff)lin}subscriptΔsubscript𝛾𝑒𝑓𝑓𝑒𝑥𝑝subscriptΔsubscript𝛾𝑒𝑓𝑓𝐺𝑆subscriptΔsubscript𝛾𝑒𝑓𝑓𝑙𝑖𝑛\{(\Delta\gamma_{eff})_{exp},\\ (\Delta\gamma_{eff})_{GS},(\Delta\gamma_{eff})_{lin}\}{ ( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_e italic_x italic_p end_POSTSUBSCRIPT , ( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_G italic_S end_POSTSUBSCRIPT , ( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_l italic_i italic_n end_POSTSUBSCRIPT } in Fig. 10b. To estimate the crack velocity at pull-off from the retraction rates used in the experiments we used the approximate relationship in Eq. (26), and this shows that a linear theory would fit the data much better (see dashed green line in Fig. 10) if we assume a rate-dependent surface energy.

Indeed, even considering that l030.0similar-to-or-equalssubscript𝑙030.0l_{0}\simeq 30.0italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≃ 30.0 nm is a more realistic estimate of the length of the fracture process zone, still we have shown that above v=100μ𝑣100𝜇v=100\;\muitalic_v = 100 italic_μm/s the linear theory largely underestimates the effective surface energy as shown in Fig. 10b. Hence, other rate-dependent causes of dissipation seems to be at play which consistently contribute to determine the overall energy to be spent for the crack to propagate.

As we have demonstrate numerically, linear theories such as PB theory, successfully estimate the dissipation happening within the bulk material, but they fail to account the rate-dependent nonlinear dissipative processes taking place within the process zone. Clearly, the assumption of constant intrinsic fracture energy and cohesive stress in the cohesive zone where large strain, high strain rate and non linear deformations (including damage) happen, is questionable as noticed by a very recent contribution by Barthel (2024). Introducing the dissipative contribution coming from the nonlinear phenomena happening within the process zone, would ultimately require additional constants to be determined from actual measurements, unless one aims at describing all the nonlinear process happening within the process zone. Given the considerable effort in the theory in characterizing the viscoelastic linear properties, these recent models are trying mostly to understand how much of the fracture energy amplification comes from the bulk dissipation and how much from the cohesive zone process rate-dependence. In this respect, the estimate we gave in Fig. 10b suggests that in our experiments at v=104μ𝑣superscript104𝜇v=10^{4}\muitalic_v = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_μm/s the nonlinear rate-dependent dissipative contribution originated within the process zone (Δγeff)GS(Δγeff)linsubscriptΔsubscript𝛾𝑒𝑓𝑓𝐺𝑆subscriptΔsubscript𝛾𝑒𝑓𝑓𝑙𝑖𝑛(\Delta\gamma_{eff})_{GS}-(\Delta\gamma_{eff})_{lin}( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_G italic_S end_POSTSUBSCRIPT - ( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_l italic_i italic_n end_POSTSUBSCRIPT equals the one coming from the dissipation in the bulk (Δγeff)linsubscriptΔsubscript𝛾𝑒𝑓𝑓𝑙𝑖𝑛(\Delta\gamma_{eff})_{lin}( roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_l italic_i italic_n end_POSTSUBSCRIPT (blue curve).

8 Conclusions

We have studied the adhesive contact between a rigid Hertzian indenter and a substrate constituted by a broad spectrum viscoelastic halfspace. For the material we have adopted a Modified Power-Law (MPL) material model, originally proposed by Williams (1964), that we have extended to provide closed-form results for the creep compliance function and for the relaxation function in time domain, and also for the complex modulus and the complex compliance in the frequency domain. Notably, the MPL model is a function of only 4 parameters, the two moduli, a characteristic exponent n𝑛nitalic_n and a characteristic time τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. In particular, by changing the exponent n𝑛nitalic_n, we have shown that it is possible to have a realistic description of a broad-band viscoelastic material, which we have demonstrated by fitting the complex modulus measured for a PDMS sample.

By using a numerical model based on the Boundary Element Method (BEM), extensive numerical studies have been performed in a wide range of the unloading rate, spanning about 8 orders of magnitude. We have shown that due to viscoelasticity, the effective surface energy can be strongly enhanced with respect to the thermodynamic surface energy, nevertheless to avoid finite size effects a certain minimum contact radius has to be reached, which we named a ”threshold contact radius” a0tsubscript𝑎0𝑡a_{0t}italic_a start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT. Our numerical simulations have shown that a0tsubscript𝑎0𝑡a_{0t}italic_a start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT is independent on the material exponent, but it depends on the pull-off enhancement that has to be reached at high unloading velocity.

Provided that finite size effects are avoided (a0>a0t)subscript𝑎0subscript𝑎0𝑡\left(a_{0}>a_{0t}\right)( italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > italic_a start_POSTSUBSCRIPT 0 italic_t end_POSTSUBSCRIPT ), the theory of Persson and Brener (2005) can be used to determine the pull-off force of the spherical indenter as a function of the crack speed at pull-off with high accuracy, but only within the assumptions of the linear theory and rate-independent fracture process zone parameters. Relating the numerical results based on a Lennard-Jones force-separation law to the theory of Persson and Brener (2005) required to define a parameter α=0.3491𝛼0.3491\alpha=0.3491italic_α = 0.3491 of order unity that relates the critical stress σcsubscript𝜎𝑐\sigma_{c}italic_σ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in PB theory to the maximum stress used in the LJ law σ0subscript𝜎0\sigma_{0}italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, which was found independent on the ratio k=E0/E𝑘subscript𝐸0subscript𝐸k=E_{0}/E_{\infty}italic_k = italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT. Adhesion experiments are usually run in displacement control, and the crack speed at pull-off is certainly not a control parameter, nevertheless we have shown that the velocity of the crack at pull-off v^^𝑣\widehat{v}over^ start_ARG italic_v end_ARG scales as v^2.887r^PB1.171similar-to-or-equals^𝑣2.887superscriptsubscript^𝑟𝑃𝐵1.171\widehat{v}\simeq 2.887\widehat{r}_{PB}^{1.171}over^ start_ARG italic_v end_ARG ≃ 2.887 over^ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_P italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1.171 end_POSTSUPERSCRIPT over more then 8 orders of magnitude, which provides an extremely simple relation to roughly estimate the pull-off force starting only from the material model parameters and the unloading rate with about ±15%plus-or-minuspercent15\pm 15\%± 15 % confidence.

Finally, by using the MPL for the viscoelastic material and the developed BEM code, we have attempted a comparison between the numerical and the experimental results, which turned out to be satisfactorily accurate up to unloading rates r=100𝑟100r=100italic_r = 100 μ𝜇\muitalic_μm/s, while for faster unloading the numerical results predict lower enhancement with respect to what is measured by our experiments. This observation turns out to be in good agreement with previous Literature results Refs. (VanDonselaar et al., 2023; Tiwari et al., 2017), where similar experiments were conducted.

A non linear description of the material behaviour must be necessarily a better description than linear, so perhaps the J𝐽Jitalic_J integral approach of Schapery (Schapery, 2023) could improve our results. However, as in classical non linear fracture mechanics, we ultimately need to measure experimentally the critical value of the fracture energy, which cannot be found reliably from other material properties, in viscoelastic adhesion even if some progress is made by the linear theories, the estimate of the bulk dissipation contribution to fracture energy enhancement is not sufficient, and, ultimately, the fracture process zone rate-dependency must be measured experimentally. Hence, at present, the measurement of the Δγeff(v)Δsubscript𝛾𝑒𝑓𝑓𝑣\Delta\gamma_{eff}(v)roman_Δ italic_γ start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ( italic_v ) curve remains the only engineering approach, resulting in the phenomenological Gent and Schultz (1972) law. Notice that if we use the measured Gent-Schultz law with a power ξ=0.41𝜉0.41\xi=0.41italic_ξ = 0.41 and assume the far field material is elastic with relaxed modulus, we can solve the adhesive contact problem using the Muller solution as corrected in Ciavarella (2021). This results in a pull-off force which doesn’t scale with the same power law of the Gent-Schulz law, but with power 0.270.270.270.27 in this case, so also the Muller solution is misleading.

Appendix A Modified power law material model

A.1 Relaxation function in time and frequency domain

Let us assume to model a viscoelastic material with a continuous distribution H(τ)𝐻𝜏H\left(\tau\right)italic_H ( italic_τ ) of relaxation times, which is the so-called material relaxation spectrum, in parallel with a Hookean spring giving the material stiffness for long time. This coincides with assuming a Wiechert model (see Fig. 11) with an infinite number of Maxwell arms. The general relation for the stress σ(t)𝜎𝑡\sigma\left(t\right)italic_σ ( italic_t ) at time t𝑡titalic_t is (Eq. (2.34) in Williams (1964))333Notice that we are using the notation according to Williams (1964). In Christensen (2012) book the relaxation spectrum is defined as [H(τ)]Christensen=[H(τ)]Williams/τsubscriptdelimited-[]𝐻𝜏𝐶𝑟𝑖𝑠𝑡𝑒𝑛𝑠𝑒𝑛subscriptdelimited-[]𝐻𝜏𝑊𝑖𝑙𝑙𝑖𝑎𝑚𝑠𝜏[H(\tau)]_{Christensen}=[H(\tau)]_{Williams}/\tau[ italic_H ( italic_τ ) ] start_POSTSUBSCRIPT italic_C italic_h italic_r italic_i italic_s italic_t italic_e italic_n italic_s italic_e italic_n end_POSTSUBSCRIPT = [ italic_H ( italic_τ ) ] start_POSTSUBSCRIPT italic_W italic_i italic_l italic_l italic_i italic_a italic_m italic_s end_POSTSUBSCRIPT / italic_τ:

σ(t)={E0+0H(τ)[ddt+1/τ]τ𝑑τddt}ε(t),𝜎𝑡subscript𝐸0superscriptsubscript0𝐻𝜏delimited-[]𝑑𝑑𝑡1𝜏𝜏differential-d𝜏𝑑𝑑𝑡𝜀𝑡\sigma\left(t\right)=\left\{E_{0}+\int_{0}^{\infty}\frac{H\left(\tau\right)}{% \left[\frac{d}{dt}+1/\tau\right]\tau}d\tau\frac{d}{dt}\right\}\varepsilon\left% (t\right)\;,italic_σ ( italic_t ) = { italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_H ( italic_τ ) end_ARG start_ARG [ divide start_ARG italic_d end_ARG start_ARG italic_d italic_t end_ARG + 1 / italic_τ ] italic_τ end_ARG italic_d italic_τ divide start_ARG italic_d end_ARG start_ARG italic_d italic_t end_ARG } italic_ε ( italic_t ) , (31)

Converting Eq. (31) in the frequency domain, we get:

σ(ω)={E0+0H(τ)𝒊ω[𝒊ωτ+1]𝑑τ}ε(ω)=E¯(ω)ε(ω),𝜎𝜔subscript𝐸0superscriptsubscript0𝐻𝜏𝒊𝜔delimited-[]𝒊𝜔𝜏1differential-d𝜏𝜀𝜔¯𝐸𝜔𝜀𝜔{\sigma}\left(\omega\right)=\left\{E_{0}+\int_{0}^{\infty}\frac{H\left(\tau% \right)\boldsymbol{i}\omega}{\left[\boldsymbol{i}\omega\tau+1\right]}d\tau% \right\}{\varepsilon}\left(\omega\right)=\overline{E}\left(\omega\right){% \varepsilon}\left(\omega\right)\;,italic_σ ( italic_ω ) = { italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_H ( italic_τ ) bold_italic_i italic_ω end_ARG start_ARG [ bold_italic_i italic_ω italic_τ + 1 ] end_ARG italic_d italic_τ } italic_ε ( italic_ω ) = over¯ start_ARG italic_E end_ARG ( italic_ω ) italic_ε ( italic_ω ) , (32)

where, 𝒊𝒊\boldsymbol{i}bold_italic_i is the imaginary unit, ω𝜔\omegaitalic_ω is the angular frequency, and, by definition, E¯(ω)=E(ω)+𝒊E′′(ω)¯𝐸𝜔superscript𝐸𝜔𝒊superscript𝐸′′𝜔\overline{E}\left(\omega\right)=E^{\prime}\left(\omega\right)+\boldsymbol{i}E^% {\prime\prime}\left(\omega\right)over¯ start_ARG italic_E end_ARG ( italic_ω ) = italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω ) + bold_italic_i italic_E start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ω ) is the complex modulus, hence:

E¯(ω)=E0+0H(τ)𝒊ω[𝒊ωτ+1]𝑑τ.¯𝐸𝜔subscript𝐸0superscriptsubscript0𝐻𝜏𝒊𝜔delimited-[]𝒊𝜔𝜏1differential-d𝜏\overline{E}\left(\omega\right)=E_{0}+\int_{0}^{\infty}\frac{H\left(\tau\right% )\boldsymbol{i}\omega}{\left[\boldsymbol{i}\omega\tau+1\right]}d\tau\;.over¯ start_ARG italic_E end_ARG ( italic_ω ) = italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_H ( italic_τ ) bold_italic_i italic_ω end_ARG start_ARG [ bold_italic_i italic_ω italic_τ + 1 ] end_ARG italic_d italic_τ . (33)

In order to fit the experimental data, one can guess a certain form for the relaxation spectrum H(τ)𝐻𝜏H\left(\tau\right)italic_H ( italic_τ ). As suggested by Williams (1964), a broad-band approximation of the response of the viscoelastic material can be obtained by adopting for the relaxation spectrum a modified power law:

H(τ)=(EE0Γ(n))(τ0τ)nexp(τ0τ),𝐻𝜏subscript𝐸subscript𝐸0Γ𝑛superscriptsubscript𝜏0𝜏𝑛subscript𝜏0𝜏H\left(\tau\right)=\left(\frac{E_{\infty}-E_{0}}{\Gamma\left(n\right)}\right)% \left(\frac{\tau_{0}}{\tau}\right)^{n}\exp\left(-\frac{\tau_{0}}{\tau}\right)\;,italic_H ( italic_τ ) = ( divide start_ARG italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG roman_Γ ( italic_n ) end_ARG ) ( divide start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_τ end_ARG ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT roman_exp ( - divide start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_τ end_ARG ) , (34)

The complex modulus is E¯(ω)=E(ω)+𝒊E′′(ω)¯𝐸𝜔superscript𝐸𝜔𝒊superscript𝐸′′𝜔\overline{E}\left(\omega\right)=E^{\prime}\left(\omega\right)+\boldsymbol{i}E^% {\prime\prime}\left(\omega\right)over¯ start_ARG italic_E end_ARG ( italic_ω ) = italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω ) + bold_italic_i italic_E start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ω ) can be written in terms of the relaxation spectrum:

E¯(ω)¯𝐸𝜔\displaystyle\overline{E}\left(\omega\right)over¯ start_ARG italic_E end_ARG ( italic_ω ) =E0+0H(τ)𝒊ω[𝒊ωτ+1]𝑑τ,absentsubscript𝐸0superscriptsubscript0𝐻𝜏𝒊𝜔delimited-[]𝒊𝜔𝜏1differential-d𝜏\displaystyle=E_{0}+\int_{0}^{\infty}\frac{H\left(\tau\right)\boldsymbol{i}% \omega}{\left[\boldsymbol{i}\omega\tau+1\right]}d\tau\;,= italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_H ( italic_τ ) bold_italic_i italic_ω end_ARG start_ARG [ bold_italic_i italic_ω italic_τ + 1 ] end_ARG italic_d italic_τ , (35)
E(ω)superscript𝐸𝜔\displaystyle E^{\prime}\left(\omega\right)italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω ) =E0+0H(τ)ω2τ[1+ω2τ2]𝑑τ,absentsubscript𝐸0superscriptsubscript0𝐻𝜏superscript𝜔2𝜏delimited-[]1superscript𝜔2superscript𝜏2differential-d𝜏\displaystyle=E_{0}+\int_{0}^{\infty}\frac{H\left(\tau\right)\omega^{2}\tau}{% \left[1+\omega^{2}\tau^{2}\right]}d\tau\;,= italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_H ( italic_τ ) italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_τ end_ARG start_ARG [ 1 + italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] end_ARG italic_d italic_τ , (36)
E′′(ω)superscript𝐸′′𝜔\displaystyle E^{\prime\prime}\left(\omega\right)italic_E start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ω ) =0H(τ)ω[1+ω2τ2]𝑑τ.absentsuperscriptsubscript0𝐻𝜏𝜔delimited-[]1superscript𝜔2superscript𝜏2differential-d𝜏\displaystyle=\int_{0}^{\infty}\frac{H\left(\tau\right)\omega}{\left[1+\omega^% {2}\tau^{2}\right]}d\tau\;.= ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_H ( italic_τ ) italic_ω end_ARG start_ARG [ 1 + italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] end_ARG italic_d italic_τ . (37)

By using Eq. (34) for the relaxation spectrum H(τ)𝐻𝜏H\left(\tau\right)italic_H ( italic_τ ) one obtains

E¯(ω)¯𝐸𝜔\displaystyle\overline{E}\left(\omega\right)over¯ start_ARG italic_E end_ARG ( italic_ω ) =E0+(EE0)𝒊ωτ0exp(𝒊ωτ0)𝐄n(𝒊ωτ0),absentsubscript𝐸0subscript𝐸subscript𝐸0𝒊𝜔subscript𝜏0𝒊𝜔subscript𝜏0subscript𝐄𝑛𝒊𝜔subscript𝜏0\displaystyle=E_{0}+\left(E_{\infty}-E_{0}\right)\boldsymbol{i}\omega\tau_{0}% \exp\left(\boldsymbol{i}\omega\tau_{0}\right)\mathbf{E}_{n}\left(\boldsymbol{i% }\omega\tau_{0}\right)\;,= italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + ( italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) bold_italic_i italic_ω italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_exp ( bold_italic_i italic_ω italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) bold_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_italic_i italic_ω italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) , (38)
E(ω)superscript𝐸𝜔\displaystyle E^{\prime}\left(\omega\right)italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω ) =E0+(EE0)Γ(n){π(τ0ω)ncos(nπ2+τ0ω)csc(nπ)++(τ0ω)2Γ(n2)pFq[1;{3n2,2n2};(τ0ω)24]},absentsubscript𝐸0subscript𝐸subscript𝐸0Γ𝑛𝜋superscriptsubscript𝜏0𝜔𝑛𝑛𝜋2subscript𝜏0𝜔𝑛𝜋superscriptsubscript𝜏0𝜔2Γsubscript𝑛2psubscriptFq13𝑛22𝑛2superscriptsubscript𝜏0𝜔24\displaystyle=E_{0}+\frac{\left(E_{\infty}-E_{0}\right)}{\Gamma\left(n\right)}% \left\{\begin{array}[c]{c}\pi\left(\tau_{0}\omega\right)^{n}\cos\left(n\frac{% \pi}{2}+\tau_{0}\omega\right)\csc\left(n\pi\right)+...\\ ...+\left(\tau_{0}\omega\right)^{2}\Gamma\left(n-2\right)_{\text{p}}\text{F}_{% \text{q}}\left[1;\left\{\frac{3-n}{2},2-\frac{n}{2}\right\};-\frac{\left(\tau_% {0}\omega\right)^{2}}{4}\right]\end{array}\right\}\;,= italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + divide start_ARG ( italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG start_ARG roman_Γ ( italic_n ) end_ARG { start_ARRAY start_ROW start_CELL italic_π ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT roman_cos ( italic_n divide start_ARG italic_π end_ARG start_ARG 2 end_ARG + italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) roman_csc ( italic_n italic_π ) + … end_CELL end_ROW start_ROW start_CELL … + ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Γ ( italic_n - 2 ) start_POSTSUBSCRIPT p end_POSTSUBSCRIPT F start_POSTSUBSCRIPT q end_POSTSUBSCRIPT [ 1 ; { divide start_ARG 3 - italic_n end_ARG start_ARG 2 end_ARG , 2 - divide start_ARG italic_n end_ARG start_ARG 2 end_ARG } ; - divide start_ARG ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG ] end_CELL end_ROW end_ARRAY } , (41)
E′′(ω)superscript𝐸′′𝜔\displaystyle E^{\prime\prime}\left(\omega\right)italic_E start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ω ) =(EE0)Γ(n){π(τ0ω)nsin(nπ2+τ0ω)csc(nπ)++(τ0ω)Γ(n1)pFq[1;{1n2,3n2};(τ0ω)24]},absentsubscript𝐸subscript𝐸0Γ𝑛𝜋superscriptsubscript𝜏0𝜔𝑛𝑛𝜋2subscript𝜏0𝜔𝑛𝜋subscript𝜏0𝜔Γsubscript𝑛1psubscriptFq11𝑛23𝑛2superscriptsubscript𝜏0𝜔24\displaystyle=\frac{\left(E_{\infty}-E_{0}\right)}{\Gamma\left(n\right)}\left% \{\begin{array}[c]{c}\pi\left(\tau_{0}\omega\right)^{n}\sin\left(n\frac{\pi}{2% }+\tau_{0}\omega\right)\csc\left(n\pi\right)+...\\ ...+\left(\tau_{0}\omega\right)\Gamma\left(n-1\right)_{\text{p}}\text{F}_{% \text{q}}\left[1;\left\{1-\frac{n}{2},\frac{3-n}{2}\right\};-\frac{\left(\tau_% {0}\omega\right)^{2}}{4}\right]\end{array}\right\}\;,= divide start_ARG ( italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG start_ARG roman_Γ ( italic_n ) end_ARG { start_ARRAY start_ROW start_CELL italic_π ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT roman_sin ( italic_n divide start_ARG italic_π end_ARG start_ARG 2 end_ARG + italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) roman_csc ( italic_n italic_π ) + … end_CELL end_ROW start_ROW start_CELL … + ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) roman_Γ ( italic_n - 1 ) start_POSTSUBSCRIPT p end_POSTSUBSCRIPT F start_POSTSUBSCRIPT q end_POSTSUBSCRIPT [ 1 ; { 1 - divide start_ARG italic_n end_ARG start_ARG 2 end_ARG , divide start_ARG 3 - italic_n end_ARG start_ARG 2 end_ARG } ; - divide start_ARG ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG ] end_CELL end_ROW end_ARRAY } , (44)

where τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the characteristic time, n>0𝑛0n>0italic_n > 0 is a characteristic exponent, ω𝜔\omegaitalic_ω is the angular frequency, E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the relaxed elastic modulus, Esubscript𝐸E_{\infty}italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT is the instantaneous elastic modulus, Fqp[a;b;z]subscriptsubscriptFqp𝑎𝑏𝑧{}_{\text{p}}\text{F}_{\text{q}}[a;b;z]start_FLOATSUBSCRIPT p end_FLOATSUBSCRIPT F start_POSTSUBSCRIPT q end_POSTSUBSCRIPT [ italic_a ; italic_b ; italic_z ] is the generalized hypergeometric function, Γ(x)Γ𝑥\Gamma\left(x\right)roman_Γ ( italic_x ) is the Euler gamma function.

Refer to caption
Figure 11: The kelvin model (left) and the Wiechert model (right) for the representation of the mechanical behaviour of a viscoelastic material.

A.2 Compliance function in time and frequency domain

Let us now consider to model a viscoelastic material with an infinite series of Voigt elements, in series with a Hookean spring giving the material stiffness for short time, the so-called Kelvin model (see Fig. 11). This coincides with assuming a continuous distribution of retardation times L(τ)𝐿𝜏L\left(\tau\right)italic_L ( italic_τ ), which is the so-called material retardation spectrum. A general relation for the deformation ε(t)𝜀𝑡\varepsilon\left(t\right)italic_ε ( italic_t ) at time t𝑡titalic_t is given by (Eq. (2.42) in Williams (1964)):

ε(t)={C+0L(τ)[ddt+1/τ]τ2𝑑τ}σ(t),𝜀𝑡subscript𝐶superscriptsubscript0𝐿𝜏delimited-[]𝑑𝑑𝑡1𝜏superscript𝜏2differential-d𝜏𝜎𝑡\varepsilon\left(t\right)=\left\{C_{\infty}+\int_{0}^{\infty}\frac{L\left(\tau% \right)}{\left[\frac{d}{dt}+1/\tau\right]\tau^{2}}d\tau\right\}\sigma\left(t% \right)\;,italic_ε ( italic_t ) = { italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_L ( italic_τ ) end_ARG start_ARG [ divide start_ARG italic_d end_ARG start_ARG italic_d italic_t end_ARG + 1 / italic_τ ] italic_τ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_d italic_τ } italic_σ ( italic_t ) , (45)

where C=1/Esubscript𝐶1subscript𝐸C_{\infty}=1/E_{\infty}italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 1 / italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT is the creep compliance in the glassy limit. Converting Eq. (45) in the frequency domain gives:

ε(ω)={C+0L(τ)[𝒊ω+1/τ]τ2𝑑τ}σ(ω)=C¯(ω)σ(ω),𝜀𝜔subscript𝐶superscriptsubscript0𝐿𝜏delimited-[]𝒊𝜔1𝜏superscript𝜏2differential-d𝜏𝜎𝜔¯𝐶𝜔𝜎𝜔{\varepsilon}\left(\omega\right)=\left\{C_{\infty}+\int_{0}^{\infty}\frac{L% \left(\tau\right)}{\left[\boldsymbol{i}\omega+1/\tau\right]\tau^{2}}d\tau% \right\}{\sigma}\left(\omega\right)=\overline{C}\left(\omega\right){\sigma}% \left(\omega\right)\;,italic_ε ( italic_ω ) = { italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_L ( italic_τ ) end_ARG start_ARG [ bold_italic_i italic_ω + 1 / italic_τ ] italic_τ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_d italic_τ } italic_σ ( italic_ω ) = over¯ start_ARG italic_C end_ARG ( italic_ω ) italic_σ ( italic_ω ) , (46)

where, 𝒊𝒊\boldsymbol{i}bold_italic_i is the imaginary unit and, by definition, C¯(ω)=C(ω)𝒊C′′(ω)¯𝐶𝜔superscript𝐶𝜔𝒊superscript𝐶′′𝜔\overline{C}\left(\omega\right)=C^{\prime}\left(\omega\right)-\boldsymbol{i}C^% {\prime\prime}\left(\omega\right)over¯ start_ARG italic_C end_ARG ( italic_ω ) = italic_C start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω ) - bold_italic_i italic_C start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ω ) is the complex compliance. Hence, we have:

C¯(ω)=C+0L(τ)[𝒊ω+1/τ]dττ2.¯𝐶𝜔subscript𝐶superscriptsubscript0𝐿𝜏delimited-[]𝒊𝜔1𝜏𝑑𝜏superscript𝜏2\overline{C}\left(\omega\right)=C_{\infty}+\int_{0}^{\infty}\frac{L\left(\tau% \right)}{\left[\boldsymbol{i}\omega+1/\tau\right]}\frac{d\tau}{\tau^{2}}\;.over¯ start_ARG italic_C end_ARG ( italic_ω ) = italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_L ( italic_τ ) end_ARG start_ARG [ bold_italic_i italic_ω + 1 / italic_τ ] end_ARG divide start_ARG italic_d italic_τ end_ARG start_ARG italic_τ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (47)

We note that to match the experimental data, a specific form for the retardation spectrum L(τ)𝐿𝜏L\left(\tau\right)italic_L ( italic_τ ) could be considered. Following Williams (1964) suggestion, a broad-band approximation of the viscoelastic material response can be achieved by using a modified power law for the retardation spectrum, such as:

L(τ)=(C0CΓ(n))(ττ0)nexp(ττ0),𝐿𝜏subscript𝐶0subscript𝐶Γ𝑛superscript𝜏subscript𝜏0𝑛𝜏subscript𝜏0L\left(\tau\right)=\left(\frac{C_{0}-C_{\infty}}{\Gamma\left(n\right)}\right)% \left(\frac{\tau}{\tau_{0}}\right)^{n}\exp\left(-\frac{\tau}{\tau_{0}}\right)\;,italic_L ( italic_τ ) = ( divide start_ARG italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG start_ARG roman_Γ ( italic_n ) end_ARG ) ( divide start_ARG italic_τ end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT roman_exp ( - divide start_ARG italic_τ end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) , (48)

The complex compliance is defined as follows:

C¯(ω)=C(ω)𝒊C′′(ω),¯𝐶𝜔superscript𝐶𝜔𝒊superscript𝐶′′𝜔\overline{C}\left(\omega\right)=C^{\prime}\left(\omega\right)-\boldsymbol{i}C^% {\prime\prime}\left(\omega\right)\,,over¯ start_ARG italic_C end_ARG ( italic_ω ) = italic_C start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω ) - bold_italic_i italic_C start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ω ) , (49)

where

C¯(ω)¯𝐶𝜔\displaystyle\overline{C}\left(\omega\right)over¯ start_ARG italic_C end_ARG ( italic_ω ) =C+0L(τ)[𝒊ω+1/τ]dττ2,absentsubscript𝐶superscriptsubscript0𝐿𝜏delimited-[]𝒊𝜔1𝜏𝑑𝜏superscript𝜏2\displaystyle=C_{\infty}+\int_{0}^{\infty}\frac{L\left(\tau\right)}{\left[% \boldsymbol{i}\omega+1/\tau\right]}\frac{d\tau}{\tau^{2}}\;,= italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_L ( italic_τ ) end_ARG start_ARG [ bold_italic_i italic_ω + 1 / italic_τ ] end_ARG divide start_ARG italic_d italic_τ end_ARG start_ARG italic_τ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (50)
C(ω)superscript𝐶𝜔\displaystyle C^{\prime}\left(\omega\right)italic_C start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω ) =C+0L(τ)[1+ω2τ2]dττ,absentsubscript𝐶superscriptsubscript0𝐿𝜏delimited-[]1superscript𝜔2superscript𝜏2𝑑𝜏𝜏\displaystyle=C_{\infty}+\int_{0}^{\infty}\frac{L\left(\tau\right)}{\left[1+% \omega^{2}\tau^{2}\right]}\frac{d\tau}{\tau}\;,= italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_L ( italic_τ ) end_ARG start_ARG [ 1 + italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] end_ARG divide start_ARG italic_d italic_τ end_ARG start_ARG italic_τ end_ARG , (51)
C′′(ω)superscript𝐶′′𝜔\displaystyle C^{\prime\prime}\left(\omega\right)italic_C start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ω ) =0L(τ)ω[1+ω2τ2]𝑑τ.absentsuperscriptsubscript0𝐿𝜏𝜔delimited-[]1superscript𝜔2superscript𝜏2differential-d𝜏\displaystyle=\int_{0}^{\infty}\frac{L\left(\tau\right)\omega}{\left[1+\omega^% {2}\tau^{2}\right]}d\tau\;.= ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_L ( italic_τ ) italic_ω end_ARG start_ARG [ 1 + italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] end_ARG italic_d italic_τ . (52)

By using Eq. (48) for the retardation spectrum L(τ)𝐿𝜏L\left(\tau\right)italic_L ( italic_τ ) one obtains

C¯(ω)¯𝐶𝜔\displaystyle\overline{C}\left(\omega\right)over¯ start_ARG italic_C end_ARG ( italic_ω ) =C+(C0C)𝒊ωτ0exp(𝒊ωτ0)𝐄n(𝒊ωτ0)absentsubscript𝐶subscript𝐶0subscript𝐶𝒊𝜔subscript𝜏0𝒊𝜔subscript𝜏0subscript𝐄𝑛𝒊𝜔subscript𝜏0\displaystyle=C_{\infty}+\frac{\left(C_{0}-C_{\infty}\right)}{\boldsymbol{i}% \omega\tau_{0}}\exp\left(-\frac{\boldsymbol{i}}{\omega\tau_{0}}\right)\mathbf{% E}_{n}\left(-\frac{\boldsymbol{i}}{\omega\tau_{0}}\right)\;= italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT + divide start_ARG ( italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ) end_ARG start_ARG bold_italic_i italic_ω italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG roman_exp ( - divide start_ARG bold_italic_i end_ARG start_ARG italic_ω italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) bold_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( - divide start_ARG bold_italic_i end_ARG start_ARG italic_ω italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) (53)
C(ω)superscript𝐶𝜔\displaystyle C^{\prime}\left(\omega\right)italic_C start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω ) =C+(C0C)(τ0ω)2nΓ(n){π(τ0ω)2cos(nπ2+1τ0ω)csc(nπ)++(τ0ω)nΓ(2+n)pFq[1;{3n2,2n2};14(τ0ω)2]}absentsubscript𝐶subscript𝐶0subscript𝐶superscriptsubscript𝜏0𝜔2𝑛Γ𝑛𝜋superscriptsubscript𝜏0𝜔2𝑛𝜋21subscript𝜏0𝜔𝑛𝜋superscriptsubscript𝜏0𝜔𝑛Γsubscript2𝑛psubscriptFq13𝑛22𝑛214superscriptsubscript𝜏0𝜔2\displaystyle=C_{\infty}+\frac{\left(C_{0}-C_{\infty}\right)\left(\tau_{0}% \omega\right)^{-2-n}}{\Gamma\left(n\right)}\left\{\begin{array}[c]{c}\pi\left(% \tau_{0}\omega\right)^{2}\cos\left(n\frac{\pi}{2}+\frac{1}{\tau_{0}\omega}% \right)\csc\left(n\pi\right)+...\\ ...+\left(\tau_{0}\omega\right)^{n}\Gamma\left(-2+n\right)_{\text{p}}\text{F}_% {\text{q}}\left[1;\left\{\frac{3-n}{2},2-\frac{n}{2}\right\};-\frac{1}{4\left(% \tau_{0}\omega\right)^{2}}\right]\end{array}\right\}= italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT + divide start_ARG ( italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ) ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) start_POSTSUPERSCRIPT - 2 - italic_n end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ ( italic_n ) end_ARG { start_ARRAY start_ROW start_CELL italic_π ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_cos ( italic_n divide start_ARG italic_π end_ARG start_ARG 2 end_ARG + divide start_ARG 1 end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω end_ARG ) roman_csc ( italic_n italic_π ) + … end_CELL end_ROW start_ROW start_CELL … + ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT roman_Γ ( - 2 + italic_n ) start_POSTSUBSCRIPT p end_POSTSUBSCRIPT F start_POSTSUBSCRIPT q end_POSTSUBSCRIPT [ 1 ; { divide start_ARG 3 - italic_n end_ARG start_ARG 2 end_ARG , 2 - divide start_ARG italic_n end_ARG start_ARG 2 end_ARG } ; - divide start_ARG 1 end_ARG start_ARG 4 ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] end_CELL end_ROW end_ARRAY } (56)
C′′(ω)superscript𝐶′′𝜔\displaystyle C^{\prime\prime}\left(\omega\right)italic_C start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ω ) =(C0C)(τ0ω)1nΓ(n){πτ0ωsin(nπ2+1τ0ω)csc(nπ)++(τ0ω)nΓ(1+n)pFq[1;{1n2,3n2};14(τ0ω)2]},absentsubscript𝐶0subscript𝐶superscriptsubscript𝜏0𝜔1𝑛Γ𝑛𝜋subscript𝜏0𝜔𝑛𝜋21subscript𝜏0𝜔𝑛𝜋superscriptsubscript𝜏0𝜔𝑛Γsubscript1𝑛psubscriptFq11𝑛23𝑛214superscriptsubscript𝜏0𝜔2\displaystyle=\frac{\left(C_{0}-C_{\infty}\right)\left(\tau_{0}\omega\right)^{% -1-n}}{\Gamma\left(n\right)}\left\{\begin{array}[c]{c}\pi\tau_{0}\omega\sin% \left(n\frac{\pi}{2}+\frac{1}{\tau_{0}\omega}\right)\csc\left(n\pi\right)+...% \\ ...+\left(\tau_{0}\omega\right)^{n}\Gamma\left(-1+n\right)_{\text{p}}\text{F}_% {\text{q}}\left[1;\left\{1-\frac{n}{2},\frac{3-n}{2}\right\};-\frac{1}{4\left(% \tau_{0}\omega\right)^{2}}\right]\end{array}\right\}\;,= divide start_ARG ( italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ) ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) start_POSTSUPERSCRIPT - 1 - italic_n end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ ( italic_n ) end_ARG { start_ARRAY start_ROW start_CELL italic_π italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω roman_sin ( italic_n divide start_ARG italic_π end_ARG start_ARG 2 end_ARG + divide start_ARG 1 end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω end_ARG ) roman_csc ( italic_n italic_π ) + … end_CELL end_ROW start_ROW start_CELL … + ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT roman_Γ ( - 1 + italic_n ) start_POSTSUBSCRIPT p end_POSTSUBSCRIPT F start_POSTSUBSCRIPT q end_POSTSUBSCRIPT [ 1 ; { 1 - divide start_ARG italic_n end_ARG start_ARG 2 end_ARG , divide start_ARG 3 - italic_n end_ARG start_ARG 2 end_ARG } ; - divide start_ARG 1 end_ARG start_ARG 4 ( italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] end_CELL end_ROW end_ARRAY } , (59)

where τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the characteristic time, n>0𝑛0n>0italic_n > 0 is a characteristic exponent, ω𝜔\omegaitalic_ω is the angular frequency, C0=1/E0subscript𝐶01subscript𝐸0C_{0}=1/E_{0}italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 / italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the relaxed compliance, C=1/Esubscript𝐶1subscript𝐸C_{\infty}=1/E_{\infty}italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 1 / italic_E start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT is the instantaneous compliance, Fqp[a;b;z]subscriptsubscriptFqp𝑎𝑏𝑧{}_{\text{p}}\text{F}_{\text{q}}[a;b;z]start_FLOATSUBSCRIPT p end_FLOATSUBSCRIPT F start_POSTSUBSCRIPT q end_POSTSUBSCRIPT [ italic_a ; italic_b ; italic_z ] is the generalized hypergeometric function, Γ(x)Γ𝑥\Gamma\left(x\right)roman_Γ ( italic_x ) is the Euler gamma function.

Appendix B Details of the PB model for the effective surface energy

According to PB theory the dimensionless effective surface energy for a MPL viscoelastic material model Γ^effsubscript^Γ𝑒𝑓𝑓\widehat{\Gamma}_{eff}over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT is obtained as

Γ^eff=[1(1k)I(n,v^,Γ^eff)]1,subscript^Γ𝑒𝑓𝑓superscriptdelimited-[]11𝑘𝐼𝑛^𝑣subscript^Γ𝑒𝑓𝑓1\widehat{\Gamma}_{eff}=\left[1-\left(1-k\right)I\left(n,\widehat{v},\widehat{% \Gamma}_{eff}\right)\right]^{-1}\,,over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT = [ 1 - ( 1 - italic_k ) italic_I ( italic_n , over^ start_ARG italic_v end_ARG , over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , (60)

where I(n,v^,Γ^eff)𝐼𝑛^𝑣subscript^Γ𝑒𝑓𝑓I(n,\widehat{v},\widehat{\Gamma}_{eff})italic_I ( italic_n , over^ start_ARG italic_v end_ARG , over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ) stands for the integral in Eq. (22), which can be evaluated in closed form as:

I(n,v^,Γ^eff)𝐼𝑛^𝑣subscript^Γeff\displaystyle I(n,\widehat{v},\widehat{\Gamma}_{\text{eff}})italic_I ( italic_n , over^ start_ARG italic_v end_ARG , over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT ) =2(32n)π(3/2n)(n1)v^{4(1+n)π(1/2+n)[Γ^eff+\displaystyle=\frac{2^{(-3-2n)}\pi^{(-3/2-n)}}{(n-1)\widehat{v}}\biggl{\{}-4^{% (1+n)}\pi^{(1/2+n)}\biggl{[}\widehat{\Gamma}_{\text{eff}}+= divide start_ARG 2 start_POSTSUPERSCRIPT ( - 3 - 2 italic_n ) end_POSTSUPERSCRIPT italic_π start_POSTSUPERSCRIPT ( - 3 / 2 - italic_n ) end_POSTSUPERSCRIPT end_ARG start_ARG ( italic_n - 1 ) over^ start_ARG italic_v end_ARG end_ARG { - 4 start_POSTSUPERSCRIPT ( 1 + italic_n ) end_POSTSUPERSCRIPT italic_π start_POSTSUPERSCRIPT ( 1 / 2 + italic_n ) end_POSTSUPERSCRIPT [ over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT + (61)
2(n1)πv^pFq[12;{12n2,1n2};(Γ^eff4πv^)2]]+\displaystyle-2(n-1)\pi\widehat{v}_{\text{p}}\text{F}_{\text{q}}\biggl{[}{-}% \frac{{1}}{2};\left\{\frac{{1}}{2}{-}\frac{{n}}{2}{,1-}\frac{{n}}{2}\right\};-% \left(\frac{\widehat{\Gamma}_{\text{eff}}}{4\pi\widehat{v}}\right)^{2}\biggr{]% }\biggr{]}+- 2 ( italic_n - 1 ) italic_π over^ start_ARG italic_v end_ARG start_POSTSUBSCRIPT p end_POSTSUBSCRIPT F start_POSTSUBSCRIPT q end_POSTSUBSCRIPT [ - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ; { divide start_ARG 1 end_ARG start_ARG 2 end_ARG - divide start_ARG italic_n end_ARG start_ARG 2 end_ARG , 1 - divide start_ARG italic_n end_ARG start_ARG 2 end_ARG } ; - ( divide start_ARG over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π over^ start_ARG italic_v end_ARG end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] ] +
+2π(Γ^effv^)n[4πv^Γ[1n2]Fqp|Reg[n12;{12,2+n2};(Γ^eff4πv^)2]+\displaystyle+2\pi\left(\frac{\widehat{\Gamma}_{\text{eff}}}{\widehat{v}}% \right)^{n}\left[4\pi\widehat{v}\Gamma\left[1-\frac{n}{2}\right]\left.{}_{% \text{p}}\text{F}_{\text{q}}\right|_{\text{Reg}}\left[\frac{{n-1}}{2};\left\{% \frac{{1}}{2}{,}\frac{{2+n}}{2}\right\};-\left(\frac{\widehat{\Gamma}_{\text{% eff}}}{4\pi\widehat{v}}\right)^{2}\right]+\right.+ 2 italic_π ( divide start_ARG over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT end_ARG start_ARG over^ start_ARG italic_v end_ARG end_ARG ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT [ 4 italic_π over^ start_ARG italic_v end_ARG roman_Γ [ 1 - divide start_ARG italic_n end_ARG start_ARG 2 end_ARG ] start_FLOATSUBSCRIPT p end_FLOATSUBSCRIPT F start_POSTSUBSCRIPT q end_POSTSUBSCRIPT | start_POSTSUBSCRIPT Reg end_POSTSUBSCRIPT [ divide start_ARG italic_n - 1 end_ARG start_ARG 2 end_ARG ; { divide start_ARG 1 end_ARG start_ARG 2 end_ARG , divide start_ARG 2 + italic_n end_ARG start_ARG 2 end_ARG } ; - ( divide start_ARG over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π over^ start_ARG italic_v end_ARG end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] +
+Γ^effΓ[32n2]Fqp|Reg[n2;{32,(3+n)2};(Γ^eff4πv^)2]]},\displaystyle+\widehat{\Gamma}_{\text{eff}}\Gamma\left[\frac{3}{2}-\frac{n}{2}% \right]\left.{}_{\text{p}}\text{F}_{\text{q}}\right|_{\text{Reg}}\biggl{[}% \frac{{n}}{2};\left\{\frac{{3}}{2}{,}\frac{{(3+n)}}{2}\right\};-\left(\frac{% \widehat{\Gamma}_{\text{eff}}}{4\pi\widehat{v}}\right)^{2}\biggr{]}\biggr{]}% \biggr{\}}\;,+ over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT roman_Γ [ divide start_ARG 3 end_ARG start_ARG 2 end_ARG - divide start_ARG italic_n end_ARG start_ARG 2 end_ARG ] start_FLOATSUBSCRIPT p end_FLOATSUBSCRIPT F start_POSTSUBSCRIPT q end_POSTSUBSCRIPT | start_POSTSUBSCRIPT Reg end_POSTSUBSCRIPT [ divide start_ARG italic_n end_ARG start_ARG 2 end_ARG ; { divide start_ARG 3 end_ARG start_ARG 2 end_ARG , divide start_ARG ( 3 + italic_n ) end_ARG start_ARG 2 end_ARG } ; - ( divide start_ARG over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π over^ start_ARG italic_v end_ARG end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] ] } , (62)

where Fqp[a,b,z]subscriptsubscriptFqp𝑎𝑏𝑧{}_{\text{p}}\text{F}_{\text{q}}[{a},b,z]start_FLOATSUBSCRIPT p end_FLOATSUBSCRIPT F start_POSTSUBSCRIPT q end_POSTSUBSCRIPT [ italic_a , italic_b , italic_z ] is the generalized hypergeometric function, Fqp|Reg[a,b,z]evaluated-atsubscriptsubscriptFqpReg𝑎𝑏𝑧\left.{}_{\text{p}}\text{F}_{\text{q}}\right|_{\text{Reg}}[{a},b,z]start_FLOATSUBSCRIPT p end_FLOATSUBSCRIPT F start_POSTSUBSCRIPT q end_POSTSUBSCRIPT | start_POSTSUBSCRIPT Reg end_POSTSUBSCRIPT [ italic_a , italic_b , italic_z ] is the regularized generalized hypergeometric function and Γ[x]Γdelimited-[]𝑥\Gamma\left[x\right]roman_Γ [ italic_x ] is the gamma function (we used Wolfram Mathematica©©{}^{\text{\textcopyright}}start_FLOATSUPERSCRIPT © end_FLOATSUPERSCRIPT for algebraic manipulation).

To ease the use of Eq. (23) we report here in Fig. (12) the quantity Γ^eff1subscript^Γ𝑒𝑓𝑓1\widehat{\Gamma}_{eff}-1over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT - 1, showing that there exists two power law regimes, the first is a linear scaling where (Γ^eff1)q1V1similar-to-or-equalssubscript^Γ𝑒𝑓𝑓1subscript𝑞1superscript𝑉1(\widehat{\Gamma}_{eff}-1)\simeq q_{1}V^{1}( over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT - 1 ) ≃ italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_V start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT, with the coefficient q1=0.0125+3.136nsubscript𝑞10.01253.136𝑛q_{1}=-0.0125+3.136nitalic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = - 0.0125 + 3.136 italic_n depending on the material exponent ”n𝑛nitalic_n”, the second instead can be written as (Γ^eff1)q2Vm2similar-to-or-equalssubscript^Γ𝑒𝑓𝑓1subscript𝑞2superscript𝑉subscript𝑚2(\widehat{\Gamma}_{eff}-1)\simeq q_{2}V^{m_{2}}( over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT - 1 ) ≃ italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_V start_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, with the {q2,m2}subscript𝑞2subscript𝑚2\{q_{2},m_{2}\}{ italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT } constants depending on n𝑛nitalic_n, which is shown in the inset of Fig. 12. Notice that, a SLS would have n1.6𝑛1.6n\approx 1.6italic_n ≈ 1.6 which provide a scaling of (Γ^eff1)V0.5proportional-tosubscript^Γ𝑒𝑓𝑓1superscript𝑉0.5(\widehat{\Gamma}_{eff}-1)\propto V^{0.5}( over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT - 1 ) ∝ italic_V start_POSTSUPERSCRIPT 0.5 end_POSTSUPERSCRIPT, while broad band materials provide a much lower exponent m2subscript𝑚2m_{2}italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT as one can see in Fig. 12 (inset).

Refer to caption
Figure 12: Fit of the enhancement of the effective surface energy Γ^eff1subscript^Γ𝑒𝑓𝑓1\widehat{\Gamma}_{eff}-1over^ start_ARG roman_Γ end_ARG start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT - 1 obtained using a MPL material model in PB theory. Two power law scaling have been identified for v^<1^𝑣1\widehat{v}<1over^ start_ARG italic_v end_ARG < 1 and for v^>1^𝑣1\widehat{v}>1over^ start_ARG italic_v end_ARG > 1, which coefficients are given in the figure and in the inset as a function of the material exponent n𝑛nitalic_n.

Acknowledgment

A.M., M.T., M.C., A.P. were partly supported by the Italian Ministry of University and Research under the Programme “Department of Excellence” Legge 232/2016 (Grant No. CUP - D93C23000100001). A.P., A.M. and M.T. were supported by the European Union (ERC-2021-STG, “Towards Future Interfaces With Tuneable Adhesion By Dynamic Excitation” - SURFACE, Project ID: 101039198, CUP: D95F22000430006). Views and opinions expressed are however those of the authors only and do not necessarily reflect those of the European Union or the European Research Council. Neither the European Union nor the granting authority can be held responsible for them. A.P. was partly supported by the European Union through the program – Next Generation EU (PRIN-2022-PNRR, ”Fighting blindness with two photon polymerization of wet adhesive, biomimetic scaffolds for neurosensory REtina-retinal Pigment epitheliAl Interface Regeneration” - REPAIR, Project ID: P2022TTZZF, CUP: D53D23018570001). Q.W. was supported by National Natural Science Foundation of China (No. 12025203).

Data availability

The dataset generated for this article will be available on Zenodo.

References

  • Afferrante and Violano (2022) Afferrante, L., Violano, G., 2022. On the effective surface energy in viscoelastic hertzian contacts. Journal of the Mechanics and Physics of Solids 158, 104669.
  • Agnelli et al. (2021) Agnelli, F., Tricarico, M., Constantinescu, A., 2021. Shape-shifting panel from 3d printed undulated ribbon lattice. Extreme Mechanics Letters 42, 101089.
  • Barthel (2024) Barthel, E., 2024. The linear viscoelastic fracture theory applies to soft solids better when they are… viscoelastic. Proceedings of the Royal Society A 480, 20230561.
  • Bonfanti et al. (2020) Bonfanti, A., Kaplan, J.L., Charras, G., Kabla, A., 2020. Fractional viscoelastic models for power-law materials. Soft Matter 16, 6002–6020.
  • Carpick et al. (1999) Carpick, R.W., Ogletree, D.F., Salmeron, M., 1999. A general equation for fitting contact area and friction vs load measurements. Journal of colloid and interface science 211, 395–400.
  • Christensen (2012) Christensen, R., 2012. Theory of viscoelasticity: an introduction. Elsevier.
  • Ciavarella (2021) Ciavarella, M., 2021. Improved muller approximate solution of the pull-off of a sphere from a viscoelastic substrate. Journal of Adhesion Science and Technology 35, 2175–2183.
  • Ciavarella et al. (2021) Ciavarella, M., Cricrì, G., McMeeking, R., 2021. A comparison of crack propagation theories in viscoelastic materials. Theoretical and applied fracture mechanics 116, 103113.
  • Dusane et al. (2023) Dusane, A.R., Lenarda, P., Paggi, M., 2023. Computational modeling of viscoelastic backsheet materials for photovoltaics. Mechanics of Materials 186, 104810.
  • Efremov et al. (2017) Efremov, Y.M., Wang, W.H., Hardy, S.D., Geahlen, R.L., Raman, A., 2017. Measuring nanoscale viscoelastic parameters of cells directly from afm force-displacement curves. Scientific reports 7, 1541.
  • Felicetti et al. (2022) Felicetti, L., Chatelet, E., Latour, A., Cornuault, P.H., Massi, F., 2022. Tactile rendering of textures by an electro-active polymer piezoelectric device: mimicking friction-induced vibrations. Biotribology 31, 100211.
  • Feng (2000) Feng, J.Q., 2000. Contact behavior of spherical elastic particles: a computational study of particle adhesion and deformations. Colloids and Surfaces A: Physicochemical and Engineering Aspects 172, 175–198.
  • Forsbach et al. (2023) Forsbach, F., Heß, M., Papangelo, A., 2023. A two-scale fem-bam approach for fingerpad friction under electroadhesion. Frontiers in mechanical engineering 8, 1074393.
  • de Gennes (1996) de Gennes, P.G., 1996. Soft adhesives. Langmuir 12, 4497–4500.
  • Gent and Lai (1994) Gent, A., Lai, S.M., 1994. Interfacial bonding, energy dissipation, and adhesion. Journal of Polymer Science Part B: Polymer Physics 32, 1543–1555.
  • Gent and Schultz (1972) Gent, A., Schultz, J., 1972. Effect of wetting liquids on the strength of adhesion of viscoelastic material. The Journal of Adhesion 3, 281–294.
  • Greenwood (1997) Greenwood, J., 1997. Adhesion of elastic spheres. Proceedings of the Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences 453, 1277–1297.
  • Greenwood (2004) Greenwood, J., 2004. The theory of viscoelastic crack propagation and healing. Journal of Physics D: Applied Physics 37, 2557.
  • Greenwood and Johnson (1981) Greenwood, J., Johnson, K., 1981. The mechanics of adhesion of viscoelastic solids. Philosophical Magazine A 43, 697–711.
  • Hosseini et al. (2021) Hosseini, A.S., Hajikarimi, P., Gandomi, M., Nejad, F.M., Gandomi, A.H., 2021. Optimized machine learning approaches for the prediction of viscoelastic behavior of modified asphalt binders. Construction and Building Materials 299, 124264.
  • Huang et al. (2004) Huang, G., Wang, B., Lu, H., 2004. Measurements of viscoelastic functions of polymers in the frequency-domain using nanoindentation. Mechanics of Time-Dependent Materials 8, 345–364.
  • Hui et al. (2022) Hui, C.Y., Zhu, B., Long, R., 2022. Steady state crack growth in viscoelastic solids: A comparative study. Journal of the Mechanics and Physics of Solids 159, 104748.
  • Johnson (1987) Johnson, K.L., 1987. Contact mechanics. Cambridge university press.
  • Johnson et al. (1971) Johnson, K.L., Kendall, K., Roberts, A., 1971. Surface energy and the contact of elastic solids. Proceedings of the royal society of London. A. mathematical and physical sciences 324, 301–313.
  • Kamperman et al. (2010) Kamperman, M., Kroner, E., Del Campo, A., McMeeking, R.M., Arzt, E., 2010. Functional adhesive surfaces with “gecko” effect: The concept of contact splitting. Advanced Engineering Materials 12, 335–348.
  • Knauss (2015) Knauss, W.G., 2015. A review of fracture in viscoelastic materials. International Journal of Fracture 196, 99–146.
  • Lake and Thomas (1967) Lake, G., Thomas, A., 1967. The strength of highly elastic materials. Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences 300, 108–119.
  • Lin et al. (2022) Lin, S., Londono, C.D., Zheng, D., Zhao, X., 2022. An extreme toughening mechanism for soft materials. Soft Matter 18, 5742–5749.
  • Linghu et al. (2023) Linghu, C., Liu, Y., Tan, Y.Y., Sing, J.H.M., Tang, Y., Zhou, A., Wang, X., Li, D., Gao, H., Hsia, K.J., 2023. Overcoming the adhesion paradox and switchability conflict on rough surfaces with shape-memory polymers. Proceedings of the National Academy of Sciences 120, e2221049120.
  • Linghu et al. (2024) Linghu, C., Liu, Y., Yang, X., Li, D., Tan, Y.Y., Mohamed, H.B.M.H., Mohammad, F.B.R., Du, Z., Su, J., Li, Y., et al., 2024. Fibrillar adhesives with unprecedented adhesion strength, switchability and scalability. National Science Review , nwae106.
  • Lorenz et al. (2013) Lorenz, B., Krick, B., Mulakaluri, N., Smolyakova, M., Dieluweit, S., Sawyer, W., Persson, B., 2013. Adhesion: role of bulk viscoelasticity and surface roughness. Journal of Physics: Condensed Matter 25, 225004.
  • Lorenz et al. (2015) Lorenz, B., Oh, Y., Nam, S., Jeon, S., Persson, B., 2015. Rubber friction on road surfaces: Experiment and theory for low sliding speeds. The Journal of chemical physics 142.
  • Maghami et al. (2024) Maghami, A., Tricarico, M., Ciavarella, M., Papangelo, A., 2024. Viscoelastic amplification of the pull-off stress in the detachment of a rigid flat punch from an adhesive soft viscoelastic layer. Engineering Fracture Mechanics 298, 109898.
  • Mandriota et al. (2024) Mandriota, C., Menga, N., Carbone, G., 2024. Adhesive contact mechanics of viscoelastic materials. International Journal of Solids and Structures 290, 112685.
  • Maugis (1992) Maugis, D., 1992. Adhesion of spheres: the jkr-dmt transition using a dugdale model. Journal of colloid and interface science 150, 243–269.
  • Mazzolai et al. (2019) Mazzolai, B., Mondini, A., Tramacere, F., Riccomi, G., Sadeghi, A., Giordano, G., Del Dottore, E., Scaccia, M., Zampato, M., Carminati, S., 2019. Octopus-inspired soft arm with suction cups for enhanced gras** tasks in confined environments. Advanced Intelligent Systems 1, 1900041.
  • Müser and Persson (2022) Müser, M.H., Persson, B.N., 2022. Crack and pull-off dynamics of adhesive, viscoelastic solids. Europhysics Letters 137, 36004.
  • Nazari et al. (2024) Nazari, R., Papangelo, A., Ciavarella, M., 2024. Friction in rolling a cylinder on or under a viscoelastic substrate with adhesion. Tribology Letters 72, 50.
  • Oliver et al. (2023) Oliver, C., Dalmas, D., Scheibert, J., 2023. Adhesion in soft contacts is minimum beyond a critical shear displacement. Journal of the Mechanics and Physics of Solids 181, 105445.
  • Papangelo and Ciavarella (2017) Papangelo, A., Ciavarella, M., 2017. A maugis–dugdale cohesive solution for adhesion of a surface with a dimple. Journal of The Royal Society Interface 14, 20160996.
  • Papangelo and Ciavarella (2020) Papangelo, A., Ciavarella, M., 2020. A numerical study on roughness-induced adhesion enhancement in a sphere with an axisymmetric sinusoidal waviness using lennard–jones interaction law. Lubricants 8, 90.
  • Papangelo and Ciavarella (2023) Papangelo, A., Ciavarella, M., 2023. Detachment of a rigid flat punch from a viscoelastic material. Tribology letters 71, 48.
  • Peng et al. (2021) Peng, B., Li, Q., Feng, X.Q., Gao, H., 2021. Effect of shear stress on adhesive contact with a generalized maugis-dugdale cohesive zone model. Journal of the Mechanics and Physics of Solids 148, 104275.
  • Persson (2017) Persson, B., 2017. Crack propagation in finite-sized viscoelastic solids with application to adhesion. Europhysics Letters 119, 18002.
  • Persson (2021) Persson, B., 2021. A simple model for viscoelastic crack propagation. The European Physical Journal E 44, 1–10.
  • Persson and Brener (2005) Persson, B., Brener, E., 2005. Crack propagation in viscoelastic solids. Physical Review E 71, 036123.
  • Petroli et al. (2022) Petroli, A., Petroli, M., Romagnoli, M., Geoghegan, M., 2022. Determination of the rate-dependent adhesion of polydimethylsiloxane using an atomic force microscope. Polymer 262, 125445.
  • Popov et al. (2010) Popov, V.L., et al., 2010. Contact mechanics and friction. Springer.
  • Pritz (2005) Pritz, T., 2005. Unbounded complex modulus of viscoelastic materials and the kramers–kronig relations. Journal of sound and vibration 279, 687–697.
  • Qi et al. (2024) Qi, Y., Li, X., Venkata, S.P., Yang, X., Sun, T.L., Hui, C.Y., Gong, J.P., Long, R., 2024. Map** deformation and dissipation during fracture of soft viscoelastic solid. Journal of the Mechanics and Physics of Solids , 105595.
  • Saharuddin et al. (2020) Saharuddin, K.D., Ariff, M.H.M., Bahiuddin, I., Mazlan, S.A., Aziz, S.A.A., Nazmi, N., Fatah, A.Y.A., Mohmad, K., 2020. Constitutive models for predicting field-dependent viscoelastic behavior of magnetorheological elastomer using machine learning. Smart Materials and Structures 29, 087001.
  • Sahli et al. (2019) Sahli, R., Pallares, G., Papangelo, A., Ciavarella, M., Ducottet, C., Ponthus, N., Scheibert, J., 2019. Shear-induced anisotropy in rough elastomer contact. Physical Review Letters 122, 214301.
  • Saulnier et al. (2004) Saulnier, F., Ondarçuhu, T., Aradian, A., Raphaël, E., 2004. Adhesion between a viscoelastic material and a solid surface. Macromolecules 37, 1067–1075.
  • Schapery (2022a) Schapery, R., 2022a. Stable and unstable viscoelastic crack growth: experimental validation of nonlinear theory for rubber. International Journal of Fracture 238, 1–15.
  • Schapery (2023) Schapery, R., 2023. Crack growth in viscoelastic media with large strains: further results and validation of nonlinear theory for rubber. International Journal of Fracture 241, 121–139.
  • Schapery (1975a) Schapery, R.A., 1975a. A theory of crack initiation and growth in viscoelastic media: I. theoretical development. International Journal of fracture 11, 141–159.
  • Schapery (1975b) Schapery, R.A., 1975b. A theory of crack initiation and growth in viscoelastic media ii. approximate methods of analysis. International Journal of Fracture 11, 369–388.
  • Schapery (1984) Schapery, R.A., 1984. Correspondence principles and a generalized j integral for large deformation and fracture analysis of viscoelastic media. International journal of fracture 25, 195–223.
  • Schapery (2022b) Schapery, R.A., 2022b. A theory of viscoelastic crack growth: revisited. International Journal of Fracture 233, 1–16.
  • Shintake et al. (2018) Shintake, J., Cacucciolo, V., Floreano, D., Shea, H., 2018. Soft robotic grippers. Advanced materials 30, 1707035.
  • Tabor (1977) Tabor, D., 1977. Surface forces and surface interactions. Journal of Colloid and Interface Science 58, 2–13.
  • Tiwari et al. (2017) Tiwari, A., Dorogin, L., Bennett, A., Schulze, K., Sawyer, W., Tahir, M., Heinrich, G., Persson, B., 2017. The effect of surface roughness and viscoelasticity on rubber adhesion. Soft matter 13, 3602–3621.
  • VanDonselaar et al. (2023) VanDonselaar, K.R., Bellido-Aguilar, D.A., Safaripour, M., Kim, H., Watkins, J.J., Crosby, A.J., Webster, D.C., Croll, A.B., 2023. Silicone elastomers and the persson-brener adhesion model. The Journal of Chemical Physics 159.
  • Violano and Afferrante (2022) Violano, G., Afferrante, L., 2022. Size effects in adhesive contacts of viscoelastic media. European Journal of Mechanics-A/Solids 96, 104665.
  • Violano et al. (2021) Violano, G., Chateauminois, A., Afferrante, L., 2021. A jkr-like solution for viscoelastic adhesive contacts. Frontiers in Mechanical Engineering 7, 664486.
  • Violano et al. (2022) Violano, G., Orlando, G., Demelio, G., Afferrante, L., 2022. Adhesion of viscoelastic media: an assessment of a recent jkr-like solution, in: IOP Conference Series: Materials Science and Engineering, IOP Publishing. p. 012038.
  • Wang et al. (2024) Wang, S.Q., Fan, Z., Gupta, C., Siavoshani, A., Smith, T., 2024. Fracture behavior of polymers in plastic and elastomeric states. Macromolecules .
  • Wayne Chen et al. (2011) Wayne Chen, W., Jane Wang, Q., Huan, Z., Luo, X., 2011. Semi-analytical viscoelastic contact modeling of polymer-based materials. ASME Journal of Tribology .
  • Williams (1964) Williams, M.L., 1964. Structural analysis of viscoelastic materials. AIAA journal 2, 785–808.
  • Williams et al. (1955) Williams, M.L., Landel, R.F., Ferry, J.D., 1955. The temperature dependence of relaxation mechanisms in amorphous polymers and other glass-forming liquids. Journal of the American Chemical society 77, 3701–3707.