This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Angle-independent optimal adhesion in plane peeling of thin elastic films at large surface roughnesses

Weilin Deng Haneesh Kesari [email protected] School of Engineering, Brown University, Providence, RI 02912
Abstract

Adhesive peeling of a thin elastic film from a substrate is a classic problem in mechanics. However, many of the investigations on this topic to date have focused on peeling from substrates with flat surfaces. In this paper, we study the problem of peeling an elastic thin film from a rigid substrate that has periodic surface undulations. We allow for contact between the detached part of the film with the substrate. We give analytical results for computing the equilibrium force given the true peeling angle, which is the angle at which the detached part of the film leaves the substrate. When there is no contact we present explicit results for computing the true peeling angle from the substrate’s profile and for determining an equilibrium state’s stability solely from the substrate’s surface curvature. The general results that we derive for the case involving contact allow us to explore the regime of peeling at large surface roughnesses. Our analysis of this regime reveals that the peel-off force can be made to become independent of the peeling direction by roughening the surface. This result is in stark contrast to results from peeling on flat surfaces, where the peel-off force strongly depends on the peeling direction. Our analysis also reveals that in the large roughness regime the peel-off force achieves its theoretical upper bound, irrespective of the other particulars of the substrate’s surface profile.

keywords:
Adhesion , Thin film , Peeling , Surface roughness
PACS:
68.35.Np, 68.35.Ct

1 Introduction

Bumpy protrusions on the surface of the lotus leaf at the small length scales have been shown to endow the leaf with the property of super-hydrophobicity (non-wettability) at the large scales [1]. This non-wetting property is the source of the lotus leaf’s acclaimed self-cleaning ability. Similarly, the periodic, small-scale, wavy undulations on the skin of some sharks are thought to reduce the skin’s frictional drag [2], as shown in Figure 1a. Such intriguing links between small-scale mechanical structures and large-scale physical properties, however, are not limited to biological systems. The small-scale topography of natural surfaces, which is generally stochastic, is often referred to as surface roughness. When separating two contacting surfaces, the surfaces’ roughness is typically thought to reduce the adhesion between them [3, 4, 5, 6]. However, there are cases in which roughness is seen to actually enhance adhesion  [7, 8]. Small-scale mechanical design can not just modulate surface properties at the large-scale but, in fact, give rise to completely new phenomenon at the large scales. For example, it was shown by Kesari et al. [9, 8, 10, 11, 12] that small-scale surface topography can give rise to the phenomenon of depth-dependent hysteresis at the large scales.

In engineering sciences, there is currently tremendous interest in creating materials with novel properties, such as materials with negative Poisson’s ratios, through the use of small-scale, intricate lattice-based mechanical designs (e.g., Figure 1b). The recent popularity of 3D printing is, presumably, primarily responsible for galvanizing such interests. The concept of modulating material properties at the large scale through incorporating mechanical designs at the small-scale—in contrast to, say, through the use of chemical or metallurgical treatments—has been of interest, in fact, for the past several decades in the composites and the mathematical homogenization communities [13, 14, 15]. However, the focus in engineering and mathematics has primarily been on modulating bulk material properties, such as elastic stiffnesses and thermal conductivities. Needless to add, surface physical properties, such as adhesion, friction, etc., are no less important than bulk properties, and, as can be gleaned from the examples previously mentioned, small-scale mechanical designs can lead to not just modulation but the emergence of completely new surface properties at the large scale. Therefore, a reason behind the engineering community’s narrow focus could be that the mathematical theories that connect large-scale surface physical properties to small-scale mechanical designs are not as well developed as the ones that connect bulk physical properties to small-scale mechanical designs. In this paper, we focus on the surface property of adhesion and present a mathematical theory that connects the large-scale force needed to peel off a thin elastic film from a rigid substrate to the substrate’s small-scale surface topography.

Adhesion is one of the most important surface physical properties. (See [11, 16, 17] for discussions on this aspect.) Adhesive peeling of a thin film from a substrate is ubiquitous in biological and engineering systems and draws growing attention due to its many applications (e.g., Figure 1c). Despite its long history, the topic of thin film peeling continues to reveal new and interesting phenomena. We review some of the recent studies on this topic in §2.

In this paper we study the mechanics of peeling in a plane (2-dimensional) problem. Specifically, we study the peeling of a thin elastic film from a rigid substrate whose surface is nominally flat with superimposed periodic undulations in a single direction. To be more precise, the substrate’s topography is, respectively, invariant and periodic in two orthogonal directions that lie in the substrate’s nominal surface plane. The invariant and periodic directions are shown marked as \mathbscre3\mathbscr{e}_{3} and \mathbscre1\mathbscr{e}_{1}, respectively, in Figure 2. The applied tractions as well as all fields in the film also do not vary in the \mathbscre3\mathbscr{e}_{3} direction. We assume that the film’s de-adherence only takes place at the peeling front, which is a straight line parallel to \mathbscre3\mathbscr{e}_{3}. We find that in the regime where the substrate’s surface roughness is large, the peel-off force becomes independent of the direction in which the film’s free end is being pulled. This result is in stark contrast to the results from the seminal analysis of Rivlin [18] and Kendall [19], who studied the classical problem of plane peeling of a thin elastic film from a flat surface. In that classical problem the peel-off force strongly depends on the direction of peeling. We also find that this angle-independent peel-off force’s magnitude, in fact, equals the maximum value that is possible for the peel-off force during the plane peeling of a thin elastic film from a rigid substrate, irrespective of the details of the substrate’s profile and the direction of peeling. The aforementioned results are especially significant considering that nowadays it is routinely possible to create highly regular small-scale topographies on engineering surfaces through the use of micro-fabrication and 3D printing technologies.

We make the following assumptions in our problem. We assume that initially the thin film is perfectly adhered to the substrate, i.e., with no gaps between it and the substrate’s surface. Over the adhered region we do not allow for (tangential) slip between the thin film and the substrate’s surface. We model adhesive interactions between the film and the substrate using the JKR theory [20], as per which during peeling the system’s energy increases linearly with the area of the region of the film-substrate interface that comes de-adhered. We consider a force-controlled peeling process. The force is applied to the thin film’s free end and its magnitude and direction can vary in a fairly general way with time, as long as the force is tensile in nature. We term the angle that the applied force makes with the direction that is parallel to \mathbscre1\mathbscr{e}_{1} and points away from adhered portion of the film the nominal peeling angle, or just the peeling angle for short (Figure 2). We take the peeling process to be quasi-static in nature and ignore all inelastic and inertial effects. We do use the notion of time, however, but only so that we can speak about the sequence of events that take place in our peeling experiment. By a peeling experiment we mean a sequence of peeling angles and force magnitudes applied to the thin film, which are infinitesimally different from each other.

We take the thin film to be composed of a linear elastic material and to be of vanishing thickness. Consequently, we assume that the thin film’s elastic strain energy only depends on its stretching deformations. Specifically, we ignore any “bending energy” in the thin film. The deformations related to stretching, however, can be of finite magnitude. Despite these and the many other assumptions that we introduce over the course of the paper, our analysis reveals important and interesting new mechanics. This is possible, we believe, because we retain the important feature of contact in our problem. That is, we allow for the detached part of the thin film to come into contact with the substrate’s surface and assume that such contact is non-adhesive and frictionless. By retaining this important feature of contact we are able to investigate the regime where the substrate’s surface roughness is large. Additionally, it is in this regime where the peel-off force becomes independent of the peeling angle and the force’s magnitude becomes equal to its maximum value.

Refer to caption
Figure 1: (a) The bonnethead shark (Sphyrna tiburo) skin surface at different body locations [2]. (b) 3D printed helmet (from [21], used without permission). (c) Picking up (i–ii) and placing (iii–iv) solid objects from one substrate to another in transfer printing [22].

The outline of the paper is as follows.

We briefly review some of recent results of the mechanics of peeling in §2. We introduce the technical details of our problem in §3.1. Specifically, we describe the kinematics of our problem in §3.1.1, and present the law that we use for modeling the thin film’s de-adherence in §3.1.2. We take a configuration in our problem to be defined by the de-adhered length, the peeling angle, and the substrate’s surface profile. The de-adhered length is defined in §3.1.1, and is roughly the width of the region on the substrate’s nominal surface in the \mathbscre1\mathbscr{e}_{1} direction from which the film has been detached. In §3.1.3 we derive the sufficiency condition, which we term the global compatibility condition, for there to be no contact between the detached part of the thin film and the substrate in a given configuration. We present results for the case in which the global compatibility condition holds, i.e., in which there is no contact, in §3.2, and for the case in which it is violated in §3.3. (When the global compatibility condition is violated the configurations that occur during a peeling experiment may or may not involve contact.)

For both cases we give explicit, analytical results for computing the equilibrium force given the true peeling angle, which is the angle at which the detached part of the film leaves the substrate (Figure 3). For the case in which there is no contact, we present explicit results for computing the true peeling angle from the substrate’s profile, and for determining an equilibrium state’s stability just from the substrate’s surface curvature. We call a configuration along with the magnitude of the force acting on it a state. For the case in which there can be contact, we present an algorithm (see Algorithm 1) that allows for numerically calculating the a configuration’s true peeling angle, and determining an equilibrium state’s stability.

For both cases, we also prove that there exist critical force values, such that if the magnitude of the applied force lies outside their range, the film either continuously de-adheres or adheres until the film is completely peeled off from or adhered to the substrate. We term these quantities the peel-initiation and peel-off forces. We determine the peel-initiation force’s minimum value as well the peel-off force’s maximum value for the wide class of surface profiles that we consider in our problem and all admissible peeling angles.

We prove that in a peeling experiment in which the peeling angle is held at a fixed value that is both acute (resp. obtuse) and violates the global compatibility condition the peel-off (resp. peel-initiation) force achieves its maximum (minimum) value, irrespective of the substrate’s profile.

Again considering experiments in which the peeling angle is held constant we show in §4 that as the surface roughness becomes large the peel-off force becomes independent of the experiment’s peeling angle. This happens irrespective of the substrate’s surface profile. The surface roughness becoming large is equivalent to the scenario where the lateral length scales in the substrate’s surface topography becomes small. Furthermore, we also show that the magnitude of that angle-independent, peel-off force equals the maximum that is possible to be achieved.

We conclude the paper by discussing the effect of the thin film’s elastic bending energy on the peel-off force in §5.

2 Literature review

The mechanics of thin film peeling has been extensively studied over the decades after the seminal work of Rivlin [18] and Kendall [19], who analyzed the peeling of a thin film from a flat surface and its importance in understanding the adhesion mechanisms in thin film-substrate systems. For example, Pesika et al. [23] presented a peel-zone model to study the effect of the peeling angle and friction on adhesion based on the microscopic observation of the geometry of the peel zone during film detachment. Chen et al. [24] considered the effect of pre-stretching the film in Kendall’s peeling model and found that the pre-tension significantly increases the peeling force at small peeling angles while decreasing it at large angles. Molinari and Ravichandran [25] proposed a general model for the peeling of non-linearly elastic thin films and investigated the effects of large deformations and pre-stretching. Begley et al. [26] developed a finite deformation analytical model for the peeling of an elastic tape and defined an effective mixed-mode interface toughness to account for frictional sliding between the surfaces in the adhered region. Gialamas et al. [27] used a Dugdale-type cohesive zone to model adhesion between an incompressible neo-Hookean elastic membrane and a flat substrate and carried out both single- and double-sided peeling analysis, while ignoring the membrane’s bending stiffness. Menga et al. [28] investigated the periodic double-sided peeling of an elastic thin film from a deformable layer that is supported by a rigid foundation. Kim and Aravas [29] performed an elasto-plastic analysis of a thin film peeling problem in which the de-adhered part of the film is in pure-bending. Kinloch et al. [30] performed an elasto-plastic analysis of the peeling of flexible laminates and calculated the fracture energy, which included not just the energy required to break the interfacial adhesive bonds but also the energy dissipated in the plastic/viscoelastic zone at the peeling front. Wei and Hutchinson [31] numerically investigated the steady-state, elasto-plastic peeling of a thin film from an elastic substrate using a cohesive zone model. Loukis et al. [32] analyzed the peeling of a viscoelastic thin film from a rigid substrate and related the fracture energy and peel-off force to the peeling speed and film thickness. Afferrante and Carbone [33] investigated the peeling of an elastic thin film from a flat viscoelastic substrate and gave a closed-form expression relating the peeling force to the peeling angle and the work of adhesion.

Spatial heterogeneity in the thin film’s elasticity and the interface’s strength can lead to significant enhancement of the effective peel-off force. The role of spatial heterogeneity in peeling was first highlighted by Kendall [34] who carried out peeling experiments using an elastic thin film that had alternating large and small bending stiffness regions. He varied the stiffness by either introducing reinforcements at select positions on a uniform thin film or by varying the film’s thickness along its length. He peeled the film from a rigid substrate at a constant force and observed abrupt changes in the speed of the peeling front at the boundaries between the stiff and compliant regions, and an overall enhancement in the peeling force. Ghatak et al. [35] and Chung and Chaudhury [36] performed displacement-controlled peeling experiments between a flexible plate and an incision-patterned thin elastic layer and found that the patterns significantly enhanced the effective adhesion. They attributed the enhancement to arrest and re-initiation of the peeling front motion at the edges of the features in the pattern. More recently, Xia et al. [37, 38, 39] experimentally and theoretically investigated the peeling of an elastic thin film while spatially varying the bending stiffness and interface strength in it. They showed that the film’s large-scale (“effective”) adhesive properties could be significantly enhanced through the use of small-scale spatial heterogenity.

There have been several studies that investigate the effect of surface roughness on thin film adhesion. For instance, Zhao et al. [40] simulated the peeling of a hyperelastic thin film from a rough substrate using finite elements and a cohesive zone model. They found that the peeling force could be increased by introducing a hierarchical wavy interface between the film and the substrate. Ghatak [41] theoretically investigated the peeling of a flexible plate from an adhesive layer which was supported on a rigid substrate and had spatially varying surface topography and elastic modulus. He found that the maximum adhesion enhancement took place when the surface height and elastic modulus varied in phase. Peng and Chen [42] studied the peeling of an elastic thin film from a rigid substrate having sinusoidally varying surface topography. They computed the peeling forces for different peeling angles and surface topography parameters and found that the maximum peeling force could be significantly enhanced by increasing the substrate’s surface roughness. To our knowledge there have not been any studies that consider contact between the de-adhered portion of the film and the substrate, or which study the stability of the equilibrium configurations, as we do in this current paper. The surface topography that we consider is fairly general and the regime we explore, namely where the lateral length scales in the substrate’s surface topography are much smaller than the other length scales in the problem, has also not been explored within the context of thin film peeling.

3 Theory of wavy peeling

Refer to caption
Figure 2: The schematic of peeling a thin elastic film from a rigid substrate with a periodic, wavy surface.

3.1 Model

3.1.1 Geometry and kinematics

Figure 2 shows the geometry of our wavy peeling problem. Let 𝔼\mathbb{E} be a three-dimensional, oriented, Hilbert space, and let the Euclidean point space \mathcal{E} be 𝔼\mathbb{E}’s principle homogeneous space. The elements of 𝔼\mathbb{E} have units of length. The physical objects in our peeling problem are contained in \mathcal{E}. The origin of \mathcal{E}, which we denote as OO, is marked in Figure 2. The vector set (\mathbscre^i)i\left(\hat{\mathbscr{e}}_{i}\right)_{i\in\mathcal{I}}, where the index set :=(1,2,3)\mathcal{I}:=(1,2,3), is an orthonormal set in 𝔼\mathbb{E}. That is, \mathbscre^i\mathbscre^j=δij\hat{\mathbscr{e}}_{i}\cdot\hat{\mathbscr{e}}_{j}=\delta_{ij}, where ii,jj\in\mathcal{I}, the symbol \cdot denotes the inner product operator between the elements of 𝔼\mathbb{E}, and δij\delta_{ij} is the Kronecker-delta symbol, which equals unity if i=ji=j and naught otherwise. A typical point XX\in\mathcal{E} is identified by its coordinates (xi)i3\left(x_{i}\right)_{i\in\mathcal{I}}\in\mathbb{R}^{3}111In our work the manner in which the information about a physical quantity’s units is stored is different from how that is usually done. We model different physical quantities as vectors belonging to different vector spaces. We store the information about a physical quantity’s units in the basis vectors we choose for that quantity’s vector space. For example, we may take 𝐞^1\hat{\mathbf{e}}_{1} to represent a motion of 1meters1\,\mathrm{meters} (or 1micron1\,\mathrm{micron}) in a certain fixed direction in 𝔼\mathbb{E}. In that case 𝐞1\mathbf{e}_{1}, as a consequence of its definition, would denote a motion of λmeters\lambda\,\mathrm{meters} (resp. λmicrons\lambda\,\mathrm{microns}) in the same direction as 𝐞^1\hat{\mathbf{e}}_{1}. Thus, in our work the components of a physical quantity with respect to the basis vectors chosen for its vector space will always be non-dimensional. For example the components of XX with respect to (𝐞i)(\mathbf{e}_{i}), namely (xi)i\left(x_{i}\right)_{i\in\mathcal{I}}, are dimensionless. that are defined such that X=O+ixi\mathbscreiX=O+\sum_{i\in\mathcal{I}}x_{i}\mathbscr{e}_{i}. The vector ixi\mathbscrei𝔼\sum_{i\in\mathcal{I}}x_{i}\mathbscr{e}_{i}\in\mathbb{E} is called XX’s position vector.

Note that the vector set (\mathbscrei)i(\mathbscr{e}_{i})_{i\in\mathcal{I}} is different from (\mathbscre^i)i(\hat{\mathbscr{e}}_{i})_{i\in\mathcal{I}}; it is defined as \mathbscrei:=λ\mathbscre^i\mathbscr{e}_{i}:=\lambda\hat{\mathbscr{e}}_{i}, where λ>0\lambda>0, for i=1,2i=1,~{}2, and \mathbscre3:=\mathbscre^3\mathbscr{e}_{3}:=\hat{\mathbscr{e}}_{3}. The set (\mathbscrei)i\left(\mathbscr{e}_{i}\right)_{i\in\mathcal{I}} is an orthogonal, but not an orthonormal, basis for 𝔼\mathbb{E}. We will be following the Einstein summation convention in this paper222We follow the Einstein summation convention in this paper. If a symbol has an italicized, light-faced latin character that appears as its subscript/superscript then that subscript/superscript denotes an index and that symbol along with that subscript/superscript denotes a component of a linear mapping. An index that appears only once in a term is called a “free” index. A free index in a term denotes that the term in fact represents the tuple of terms created by varying the free index in the original term over \mathcal{I}. An index that appears twice in a term is called a “repeated index”. A repeated index in a term denotes a sum of the terms that are created by varying the free index in the original term over \mathcal{I}., and write expressions such as (\mathbscrei)i\left(\mathbscr{e}_{i}\right)_{i\in\mathcal{I}} and ixi\mathbscrei\sum_{i\in\mathcal{I}}x_{i}\mathbscr{e}_{i} simply as (\mathbscrei)\left(\mathbscr{e}_{i}\right) and xi\mathbscreix_{i}\mathbscr{e}_{i}, respectively, and lists such as \mathbscre1,\mathbscre2\mathbscr{e}_{1},~{}\mathbscr{e}_{2}, and \mathbscre3\mathbscr{e}_{3} simply as \mathbscrei\mathbscr{e}_{i}.

The substrate is a rigid solid whose points’ position vectors belong to the set

𝕊={x1\mathbscre1+x2\mathbscre2+x3\mathbscre3𝔼|(x1,x2,x3)3andx2<ρ(x1):=αϱ(x1)},\mathbb{S}=\left\{x_{1}\mathbscr{e}_{1}+x_{2}\mathbscr{e}_{2}+x_{3}\mathbscr{e}_{3}\in\mathbb{E}~{}|~{}(x_{1},x_{2},x_{3})\in\mathbb{R}^{3}~{}\text{and}~{}x_{2}<\rho(x_{1}):=\alpha\varrho(x_{1})\right\}, (3.1)

where α:=A/λ\alpha:=A/\lambda, A0A\geq 0, ϱ:[1,1]\varrho:\mathbb{R}\to[-1,1] is a surjective, 1-periodic function, and \mathbb{R} is the set of real numbers. By 1-periodic we mean that ϱ(x1+1)=ϱ(x1)\varrho(x_{1}+1)=\varrho(x_{1}) for all x1x_{1}\in\mathbb{R}. Without loss of generality we take 01ϱ(x1)𝑑x1=0\int_{0}^{1}\varrho(x_{1})\,dx_{1}=0. We also assume that ϱ\varrho and its first and second derivatives, which we denote as ϱ˙\dot{\varrho} and ϱ¨\ddot{\varrho}, respectively, are non-constant, continuous functions, i.e., ϱC2\varrho\in C^{2}. We call α\alpha, AA, λ\lambda, and ϱ\varrho the substrate surface’s aspect ratio, amplitude, periodicity, and profile, respectively.

We take the thin film to be of width 𝒷\mathscr{b}, of thickness 𝒽\mathscr{h}, and to be perfectly adhered to the substrate’s surface in its initial, or reference, configuration. Here 𝒷\mathscr{b} and 𝒽\mathscr{h} are physical parameters and have units of length. Say the units of \mathbscrei\mathbscr{e}_{i} and \mathbscre^i\hat{\mathbscr{e}}_{i} is some length 𝓂\mathscr{m}, which, for example, may stand for meters, then we would say that the magnitude of 𝒷\mathscr{b} (resp. 𝒽\mathscr{h}) is bb (resp. hh) iff 𝒷=b𝓂\mathscr{b}=b\,\mathscr{m} (resp. 𝒽=h𝓂\mathscr{h}=h\,\mathscr{m}). Thus, initially the configuration of the thin film can be described as

𝔹0={x1\mathbscre1+ρ(x1)\mathbscre2+x3\mathbscre3𝔼|x1𝒟,|x3|b/2},\mathbb{B}_{0}=\left\{x_{1}\mathbscr{e}_{1}+\rho\left(x_{1}\right)\mathbscr{e}_{2}+x_{3}\mathbscr{e}_{3}\in\mathbb{E}\;|\;x_{1}\in\mathcal{D},\;|x_{3}\rvert\leq b/2\right\}, (3.2)

where 𝒟:=[0,+)\mathcal{D}:=[0,+\infty)333As is standard in continuum mechanics, we will be referring to a particular thin film material particle using the coordinates of the the spatial point that it occupied in the initial configuration. That is, when we say a material particle (x1,x2,x3)(x_{1},x_{2},x_{3}), we in fact are referring to the material particle that in the initial configuration occupied the spatial point with coordinates (x1,x2,x3)(x_{1},x_{2},x_{3}). For the sake of brevity, we will be referring to a material particle (x1,x2,x3)(x_{1},x_{2},x_{3}) simply as (x1,x3)(x_{1},x_{3}), since the second co-ordinate of any spatial point that was occupied by a thin film material particle in the initial configuration is fully determined by the point’s first co-ordinate (cf. (3.2)). Finally, when we say “the (thin-film) material particle x1x_{1}”, where x1𝒟x_{1}\in\mathcal{D}, we in fact mean the group of material particles (x1,x3)(x_{1},x_{3}), where |x3|b/2\lvert x_{3}\rvert\leq b/2. .

We model the peeling process by assuming that any deformed configuration 𝜿\bm{\kappa} of the thin film can be described as444In standard continuum mechanics the reference and deformed configurations are taken to belong to different spaces. In contrast, here we take both the reference and deformed configurations, 𝔹0\mathbb{B}_{0} and 𝔹\mathbb{B}, to belong to the space space, namely 𝔼\mathbb{E}.

𝔹={\mathbscrx(x1)+x3\mathbscre3𝔼|x1𝒟,|x3|b/2},\displaystyle\mathbb{B}=\left\{\mathbscr{x}(x_{1})+x_{3}\mathbscr{e}_{3}\in\mathbb{E}\;|\;x_{1}\in\mathcal{D},\;|x_{3}\rvert\leq b/2\right\}, (3.3a)
where
\mathbscrx(x1):=(x1+u1(x1))\mathbscre1+(ρ(x1)+u2(x1))\mathbscre2,\displaystyle\mathbscr{x}(x_{1}):=\left(x_{1}+u_{1}(x_{1})\right)\mathbscr{e}_{1}+\left(\rho(x_{1})+u_{2}(x_{1})\right)\mathbscr{e}_{2}, (3.3b)

u1,u2:𝒟u_{1},u_{2}:\mathcal{D}\to\mathbb{R} and their derivatives, denoted as u˙1\dot{u}_{1} and u˙2\dot{u}_{2}, are continuous. We further assume that u1(x1)=u2(x1)=0u_{1}\left(x_{1}\right)=u_{2}\left(x_{1}\right)=0 for all x1ax_{1}\geq a, for some a𝒟a\in\mathcal{D}, and u˙1(x1)=u˙2(x1)=0\dot{u}_{1}\left(x_{1}\right)=\dot{u}_{2}\left(x_{1}\right)=0 for all x1>ax_{1}>a. We refer to aa as the de-adhered length. We call P:={O+a\mathbscre1+ρ(a)\mathbscre2+x3\mathbscre3||x3|b/2}P:=\left\{O+a\mathbscr{e}_{1}+\rho(a)\mathbscr{e}_{2}+x_{3}\mathbscr{e}_{3}\in\mathcal{E}\;|\;\lvert x_{3}\rvert\leq b/2\right\} the peeling front and Γa={x1𝒟|x1>a}\Gamma_{a}=\{x_{1}\in\mathcal{D}~{}|~{}x_{1}>a\} the adhered region.

On knowing the de-adhered length aa the length of the film peeled from the substrate can be computed as λl(a;ρ)𝓂\lambda l(a\,;\rho)\,\mathscr{m}, where l(;ρ):𝒟𝒟l(\cdot\,;\rho):\mathcal{D}\to\mathcal{D} is defined by the equation

l(a;ρ)=0a𝑑x1(1+ρ˙(x1)2)1/2.l(a\,;\rho)=\int_{0}^{a}dx_{1}\left(1+\dot{\rho}(x_{1})^{2}\right)^{1/2}. (3.4)

We refer to l(a;ρ)l(a\,;\rho) as the peeled length, and will often abbreviate it as ll.

3.1.2 Evolution law for de-adherence

In postulating the evolution law for de-adherence, we ignore all dynamic effects, such as inertial forces, kinetic energy, and viscoelastic behavior in the thin film. As we mentioned in §1, we do use the notion of time, but only so that we can speak about the sequence of events that take place in our peeling experiment.

In our peeling experiment the thin film de-adheres due to the application of the force 𝖋\bm{\mathfrak{f}} to its free end, which initially is at OO, the origin of \mathcal{E}. We think of force as a linear map from 𝔼\mathbb{E} into a one dimensional vector space whose elements have units of energy. The set of all forces can, of course, be made into a vector space, which we will denote as 𝔽\mathbb{F}. We use the set (𝖋i)\left(\bm{\mathfrak{f}}_{i}\right) as a basis for 𝔽\mathbb{F}, where 𝖋i\bm{\mathfrak{f}}_{i} are defined such that 𝖋i(\mathbscrej)=Ecδij\bm{\mathfrak{f}}_{i}\left(\mathbscr{e}_{j}\right)=\mathbio{E}_{c}\,\delta_{ij}. The symbol Ec\mathbio{E}_{c}, which appears in the last expression, is defined to be equal to the energy λ𝒷𝒽𝓂\lambda{\mathpzc{E}}\mathscr{b}\mathscr{h}\mathscr{m}, where {\mathpzc{E}} is the thin film’s Young’s modulus. We express 𝖋\bm{\mathfrak{f}} as F𝖋^F\hat{\bm{\mathfrak{f}}}, where F0F\geq 0 and 𝖋^:=cos(θ)𝖋1+sin(θ)𝖋2\hat{\bm{\mathfrak{f}}}:=-\cos(\theta)\bm{\mathfrak{f}}_{1}+\sin(\theta)\bm{\mathfrak{f}}_{2}. The angle θ\theta is shown marked in Figure 3. It is prescribed as part of the problem’s definition. We call it the nominal peeling angle, or simply peeling angle for short. We denote the position vector of the free end of the thin film, on which 𝖋\bm{\mathfrak{f}} acts, as \mathbscru𝔼\mathbscr{u}\in\mathbb{E}, and will often refer to this quantity as simply the force-position-vector. Our experiment is force controlled and quasi-static. By force-controlled we mean that while the system’s configuration is changing, 𝖋\bm{\mathfrak{f}} is held fixed. If our experiment were a general force controlled experiment, then we would be allowed to vary 𝖋\bm{\mathfrak{f}} once the adhered region had stopped evolving. However, in our experiment we are only allowed to vary FF after the adhered region has stopped evolving. By quasi-static we mean that each variation in FF is of infinitesimal magnitude and is applied all at once at a particular instance in time.

Consider an experimentally observed configuration of the thin film 𝜿0\bm{\kappa}_{0}^{\leavevmode\hbox to1.9pt{\vbox to1.9pt{\pgfpicture\makeatletter\hbox{\hskip 0.95pt\lower-0.95pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{}{{{}} {}{}{}{}{}{}{}{} }\pgfsys@beginscope\pgfsys@invoke{ }\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{0.75pt}{0.0pt}\pgfsys@curveto{0.75pt}{0.41422pt}{0.41422pt}{0.75pt}{0.0pt}{0.75pt}\pgfsys@curveto{-0.41422pt}{0.75pt}{-0.75pt}{0.41422pt}{-0.75pt}{0.0pt}\pgfsys@curveto{-0.75pt}{-0.41422pt}{-0.41422pt}{-0.75pt}{0.0pt}{-0.75pt}\pgfsys@curveto{0.41422pt}{-0.75pt}{0.75pt}{-0.41422pt}{0.75pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}, in which the de-adhered length is aΔaa-\Delta a, the peeled length is lΔll-\Delta l, the force position vector is \mathbscruΔ\mathbscru\mathbscr{u}-\Delta\mathbscr{u}, and the force is (FΔF)𝖋^(F-\Delta F)\hat{\bm{\mathfrak{f}}}. Imagine that a force variation of ΔF𝖋^\Delta F\hat{\bm{\mathfrak{f}}} is then abruptly added to the force that was previously acting on the thin film so that the new force is F𝖋^F\hat{\bm{\mathfrak{f}}}. As a result, the thin film’s configuration will evolve to a new configuration 𝜿\bm{\kappa}^{\leavevmode\hbox to1.9pt{\vbox to1.9pt{\pgfpicture\makeatletter\hbox{\hskip 0.95pt\lower-0.95pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{}{{{}} {}{}{}{}{}{}{}{} }\pgfsys@beginscope\pgfsys@invoke{ }\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{0.75pt}{0.0pt}\pgfsys@curveto{0.75pt}{0.41422pt}{0.41422pt}{0.75pt}{0.0pt}{0.75pt}\pgfsys@curveto{-0.41422pt}{0.75pt}{-0.75pt}{0.41422pt}{-0.75pt}{0.0pt}\pgfsys@curveto{-0.75pt}{-0.41422pt}{-0.41422pt}{-0.75pt}{0.0pt}{-0.75pt}\pgfsys@curveto{0.41422pt}{-0.75pt}{0.75pt}{-0.41422pt}{0.75pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}, in which the de-adhered length is aa, the peeled length is ll, the force-position-vector is \mathbscru\mathbscr{u}, and, of course, the force is F𝖋^F\hat{\bm{\mathfrak{f}}}. We postulate that the new configuration 𝜿\bm{\kappa}^{\leavevmode\hbox to1.9pt{\vbox to1.9pt{\pgfpicture\makeatletter\hbox{\hskip 0.95pt\lower-0.95pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{}{{{}} {}{}{}{}{}{}{}{} }\pgfsys@beginscope\pgfsys@invoke{ }\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{0.75pt}{0.0pt}\pgfsys@curveto{0.75pt}{0.41422pt}{0.41422pt}{0.75pt}{0.0pt}{0.75pt}\pgfsys@curveto{-0.41422pt}{0.75pt}{-0.75pt}{0.41422pt}{-0.75pt}{0.0pt}\pgfsys@curveto{-0.75pt}{-0.41422pt}{-0.41422pt}{-0.75pt}{0.0pt}{-0.75pt}\pgfsys@curveto{0.41422pt}{-0.75pt}{0.75pt}{-0.41422pt}{0.75pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}} is the one that locally minimizes the system’s total potential energy and is closest to 𝜿0\bm{\kappa}_{0}^{\leavevmode\hbox to1.9pt{\vbox to1.9pt{\pgfpicture\makeatletter\hbox{\hskip 0.95pt\lower-0.95pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{}{{{}} {}{}{}{}{}{}{}{} }\pgfsys@beginscope\pgfsys@invoke{ }\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{0.75pt}{0.0pt}\pgfsys@curveto{0.75pt}{0.41422pt}{0.41422pt}{0.75pt}{0.0pt}{0.75pt}\pgfsys@curveto{-0.41422pt}{0.75pt}{-0.75pt}{0.41422pt}{-0.75pt}{0.0pt}\pgfsys@curveto{-0.75pt}{-0.41422pt}{-0.41422pt}{-0.75pt}{0.0pt}{-0.75pt}\pgfsys@curveto{0.41422pt}{-0.75pt}{0.75pt}{-0.41422pt}{0.75pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}.

To make the postulate precise, consider a configuration 𝜿~\widetilde{\bm{\kappa}} that is close to 𝜿\bm{\kappa}^{\leavevmode\hbox to1.9pt{\vbox to1.9pt{\pgfpicture\makeatletter\hbox{\hskip 0.95pt\lower-0.95pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{}{{{}} {}{}{}{}{}{}{}{} }\pgfsys@beginscope\pgfsys@invoke{ }\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{0.75pt}{0.0pt}\pgfsys@curveto{0.75pt}{0.41422pt}{0.41422pt}{0.75pt}{0.0pt}{0.75pt}\pgfsys@curveto{-0.41422pt}{0.75pt}{-0.75pt}{0.41422pt}{-0.75pt}{0.0pt}\pgfsys@curveto{-0.75pt}{-0.41422pt}{-0.41422pt}{-0.75pt}{0.0pt}{-0.75pt}\pgfsys@curveto{0.41422pt}{-0.75pt}{0.75pt}{-0.41422pt}{0.75pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}} and has the same force as 𝜿\bm{\kappa}^{\leavevmode\hbox to1.9pt{\vbox to1.9pt{\pgfpicture\makeatletter\hbox{\hskip 0.95pt\lower-0.95pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{}{{{}} {}{}{}{}{}{}{}{} }\pgfsys@beginscope\pgfsys@invoke{ }\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{0.75pt}{0.0pt}\pgfsys@curveto{0.75pt}{0.41422pt}{0.41422pt}{0.75pt}{0.0pt}{0.75pt}\pgfsys@curveto{-0.41422pt}{0.75pt}{-0.75pt}{0.41422pt}{-0.75pt}{0.0pt}\pgfsys@curveto{-0.75pt}{-0.41422pt}{-0.41422pt}{-0.75pt}{0.0pt}{-0.75pt}\pgfsys@curveto{0.41422pt}{-0.75pt}{0.75pt}{-0.41422pt}{0.75pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}, i.e., F𝖋^F\hat{\bm{\mathfrak{f}}}. The de-adhered and peeled lengths, and the force-position-vector in 𝜿~\widetilde{\bm{\kappa}} are, however, different from those in 𝜿\bm{\kappa}^{\leavevmode\hbox to1.9pt{\vbox to1.9pt{\pgfpicture\makeatletter\hbox{\hskip 0.95pt\lower-0.95pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{}{{{}} {}{}{}{}{}{}{}{} }\pgfsys@beginscope\pgfsys@invoke{ }\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{0.75pt}{0.0pt}\pgfsys@curveto{0.75pt}{0.41422pt}{0.41422pt}{0.75pt}{0.0pt}{0.75pt}\pgfsys@curveto{-0.41422pt}{0.75pt}{-0.75pt}{0.41422pt}{-0.75pt}{0.0pt}\pgfsys@curveto{-0.75pt}{-0.41422pt}{-0.41422pt}{-0.75pt}{0.0pt}{-0.75pt}\pgfsys@curveto{0.41422pt}{-0.75pt}{0.75pt}{-0.41422pt}{0.75pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}; we denote them, respectively, as a+δaa+\delta a, l+δll+\delta l, and \mathbscru+δ\mathbscru\mathbscr{u}+\delta\mathbscr{u}. Since F0F\geq 0, i.e. the peeled section of the film is in tension, the variations δl\delta l and δ\mathbscru\delta\mathbscr{u}, in fact, depend on δa\delta a. Thus, these variations are to be interpreted as abbreviations for δl(δa;a,F,ρ)\delta l(\delta a\,;a,F,\rho) and δ\mathbscru(δa;a,F,ρ)\delta\mathbscr{u}(\delta a\,;a,F,\rho), respectively. Let the difference in the system’s potential energy between the configurations 𝜿~\widetilde{\bm{\kappa}} and 𝜿\bm{\kappa}^{\leavevmode\hbox to1.9pt{\vbox to1.9pt{\pgfpicture\makeatletter\hbox{\hskip 0.95pt\lower-0.95pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{}{{{}} {}{}{}{}{}{}{}{} }\pgfsys@beginscope\pgfsys@invoke{ }\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{0.75pt}{0.0pt}\pgfsys@curveto{0.75pt}{0.41422pt}{0.41422pt}{0.75pt}{0.0pt}{0.75pt}\pgfsys@curveto{-0.41422pt}{0.75pt}{-0.75pt}{0.41422pt}{-0.75pt}{0.0pt}\pgfsys@curveto{-0.75pt}{-0.41422pt}{-0.41422pt}{-0.75pt}{0.0pt}{-0.75pt}\pgfsys@curveto{0.41422pt}{-0.75pt}{0.75pt}{-0.41422pt}{0.75pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}} be EcδE\mathbio{E}_{c}\,\delta E, where Ec\mathbio{E}_{c} is a constant and has unit of energy, and we refer to δE\delta E\in\mathbb{R} as the non-dimensional potential energy variation. In our model of the wavy peeling experiment, we take δE\delta E to be a sum of three different energy variations. These variations take place, respectively, in the energy stored in the interbody adhesive interactions between the thin film and the substrate, the elastic strain energy stored in the peeled part of the thin film, and, finally, the energy stored in the apparatus that maintains the constant force F𝖋^F\hat{\bm{\mathfrak{f}}} between the configuration 𝜿\bm{\kappa}^{\leavevmode\hbox to1.9pt{\vbox to1.9pt{\pgfpicture\makeatletter\hbox{\hskip 0.95pt\lower-0.95pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{}{{{}} {}{}{}{}{}{}{}{} }\pgfsys@beginscope\pgfsys@invoke{ }\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{0.75pt}{0.0pt}\pgfsys@curveto{0.75pt}{0.41422pt}{0.41422pt}{0.75pt}{0.0pt}{0.75pt}\pgfsys@curveto{-0.41422pt}{0.75pt}{-0.75pt}{0.41422pt}{-0.75pt}{0.0pt}\pgfsys@curveto{-0.75pt}{-0.41422pt}{-0.41422pt}{-0.75pt}{0.0pt}{-0.75pt}\pgfsys@curveto{0.41422pt}{-0.75pt}{0.75pt}{-0.41422pt}{0.75pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}} and 𝜿~\widetilde{\bm{\kappa}}.

We model adhesive interactions between the thin film and the substrate using the JKR theory [20]. According to this theory, the formation of an interface region of area δ𝒜\delta\mathscr{A} lowers the system’s total potential energy, irrespective of the shape of the interface region or any other details of the experiment, reversibly by 𝓌δ𝒜\mathscr{w}\delta\mathscr{A}. Here, 𝓌\mathscr{w} is the Dupré work of adhesion [43, p. 30]. Thus, the variation in the potential energy stored in the interbody adhesive interactions is

wδl,w\delta l, (3.5)

where ww is defined such that 𝒽w=𝓌{\mathpzc{E}}\mathscr{h}w=\mathscr{w}. That is, ww is the non-dimensional work of adhesion.

We assume that the peeled part of the film is in uniform, uniaxial tension. It follows from this assumption that the strain in the peeled part of the film is also uniform, and uniaxial, and that its magnitude is equal to FF. It also follows that the variation in the elastic strain energy stored in the thin film is

12F2δl.\frac{1}{2}F^{2}\delta l. (3.6)

Usually, the above term is supplemented by an additional term that corresponds to bending energy in the thin film [42]. We, however, ignore the film’s bending energy in our model. As we state in our closing remarks, we believe that ignoring the bending energy is unlikely to significantly affect our estimates for the film’s peel-off force in the type of the peeling experiments that we consider in this paper.

The variation in the potential energy of the apparatus maintaining the constant force F𝖋^F\hat{\bm{\mathfrak{f}}} is 𝖋(δ\mathbscru)-\bm{\mathfrak{f}}(\delta\mathbscr{u}). Say δ\mathbscru=δuj\mathbscrej\delta\mathbscr{u}=\delta u_{j}\mathbscr{e}_{j}. This variation can therefore be written as

𝑭δ𝒖,-\bm{F}\cdot\delta\bm{u}, (3.7)

where 𝑭:=F𝑭^\bm{F}:=F\hat{\bm{F}}, 𝑭^:=(cos(θ),sin(θ),0)\hat{\bm{F}}:=(-\cos(\theta),\sin(\theta),0), and δ𝒖:=(δui)\delta\bm{u}:=\left(\delta u_{i}\right). It follows from the previous discussion that,

δE=wδl+12F2δl𝑭δ𝒖.\delta E=w\delta l+\frac{1}{2}F^{2}\delta l-\bm{F}\cdot\delta\bm{u}. (3.8)

As is the case with δl\delta l and δ𝒖\delta\bm{u}, the symbol δE\delta E in (3.8) is, in fact, an abbreviation for the value δE(δa;a,F,ρ)\delta E(\delta a;\,a,F,\rho).

Since δl\delta l and δ𝒖\delta\bm{u} depend on δa\delta a, the formula for δE\delta E given by (3.8) needs to be further refined before it can be used for determining whether or not a particular de-adhered length aa is in equilibrium; and, if aa is in equilibrium, then for determining the stability of that equilibrium. The particulars of the requisite refinement depend on whether or not the peeled part of the film contacts the substrate. Therefore, we will be making separate refinements of (3.8) for the cases of no-contact and contact in §3.2 and §3.3, respectively. In both cases, however, the refinement process will involve determining the asymptotic dependence of δl\delta l and δ𝒖\delta\bm{u} on δa\delta a as δa0\delta a\to 0, and then using that information to determine the asymptotic dependence of δE\delta E on δa\delta a.

3.1.3 Conditions for the peeled part of thin film to not contact the substrate

Let 𝑻(x1)\bm{T}\left(x_{1}\right) be (\mathbscre1\mathbscr𝒙˙(x1),\mathbscre2\mathbscr𝒙˙(x1),0)3(\mathbscr{e}_{1}\cdot\dot{\bm{\mathbscr{x}}}(x_{1}),\mathbscr{e}_{2}\cdot\dot{\bm{\mathbscr{x}}}(x_{1}),0)\in\mathbb{R}^{3}, where \mathbscrx(x1)\mathbscr{x}(x_{1}) is defined in (3.3b). Because a material particle x1x_{1} is de-adhered from or adhered to the substrate (depending on whether x1x_{1} is less than or greater than aa), the vector 𝑻(x1)\bm{T}(x_{1}) will often be discontinuous at x1=ax_{1}=a. However, since we have assumed ϱ˙\dot{\varrho}, u˙1\dot{u}_{1}, and u˙2\dot{u}_{2} to be continuous, the following right and left hand limits of 𝑻(x1)\bm{T}\left(x_{1}\right) as x1ax_{1}\to a are well defined:

𝑻(a±):=limx1a±𝑻(x1).\bm{T}\left(a^{\pm}\right):=\lim_{x_{1}\to a^{\pm}}\bm{T}(x_{1}).

The vector 𝑻(a+)\bm{T}\left(a^{+}\right) is essentially the tangent to the substrate’s surface profile at x1=ax_{1}=a, and the vector 𝑻(a)-\bm{T}\left(a^{-}\right) points in the direction the detached part of the thin film leaves the substrate’s surface at the peeling front. (See Figure 3.) We call the angle between 𝑻(a)\bm{T}\left(a^{-}\right) and 𝑻(a+)\bm{T}\left(a^{+}\right) the true peeling angle ψ(a)\psi(a). For the thin film to not go through the substrate immediately after de-adhering from it, it is necessary that

ψ(a)[0,π].\psi(a)\in[0,\pi]. (3.9)

We refer to (3.9) as the local compatibility condition.

It can be shown that once the local compatibility condition is satisfied, the peeled part of the thin film will not contact the substrate anywhere, irrespective of the location of the peeling front iff

θ\displaystyle\theta [tan1(ρ˙),πtan1(ρ˙+)],\displaystyle\in\left[-\tan^{-1}\left(\dot{\rho}^{-}\right),\pi-\tan^{-1}\left(\dot{\rho}^{+}\right)\right], (3.10a)
where
ρ˙±\displaystyle\dot{\rho}^{\pm} :=ρ˙(a±),\displaystyle:=\dot{\rho}\left(a^{\pm}\right), (3.10b)
a±\displaystyle a^{\pm} :=argmax{±ρ˙(x1)|x1[0,1]}.\displaystyle:=\arg\max\left\{\pm\,\dot{\rho}\left(x_{1}\right)~{}|~{}x_{1}\in[0,1]\right\}. (3.10c)

We refer to (3.10) as the global compatibility condition.

3.2 Peeling with no contact

Refer to caption
Figure 3: The schematic of peeling without contact during an infinitesimal advance of the peeling front from PP to P~\tilde{P}, where δ𝒖\delta\bm{u} denotes the variation of position vector of the free end of the thin film, ψ(a)\psi(a) is the true peeling angle.

In this section we study the case in which the local and global compatibility conditions (i.e., (3.9) and (3.10)) are satisfied, and therefore the peeled part of the thin film does not contact the substrate anywhere.

3.2.1 Energy variation

Since the peeled part of the thin film is not in contact with the substrate, it follows that

δ𝒖=(δa,δρ)+δl(1+ε)𝑭^,\delta{\bm{u}}=(\delta{a},\delta\rho)+\delta{l}(1+\varepsilon)\hat{\bm{F}}, (3.11)

where ε\varepsilon is the uniform, uni-axial strain in the peeled part of the thin film, and δρ:=ρ(a+δa)ρ(a)\delta\rho:={\rho}({a}+\delta{a})-{\rho}({a}). It follows from (3.4) that

δl(δa;a,ρ)=l˙(a;ρ)δa+12l¨(a;ρ)(δa)2+o((δa)2),\delta l(\delta a;a,\rho)=\dot{l}(a;\rho)\delta a+\frac{1}{2}\ddot{l}(a;\rho)\left(\delta a\right)^{2}+o\left(\left(\delta a\right)^{2}\right), (3.12)

where

l˙(a;ρ)\displaystyle\dot{l}(a;\rho) =(1+ρ˙(a)2)12,\displaystyle=\left(1+\dot{\rho}(a)^{2}\right)^{\frac{1}{2}}, (3.13a)
l¨(a;ρ)\displaystyle\ddot{l}(a;\rho) =ρ˙(a)ρ¨(a)(1+ρ˙(a)2)12.\displaystyle=\frac{\dot{\rho}(a)\ddot{\rho}(a)}{\left(1+\dot{\rho}(a)^{2}\right)^{\frac{1}{2}}}.~ (3.13b)

Substituting the asymptotic expansions for δρ\delta\rho and δl\delta l as δa0\delta a\to 0 into (3.11), and then substituting the resulting asymptotic expansion for δ𝒖\delta\bm{u} into (3.7), we find that

𝑭δ𝒖\displaystyle-{\bm{F}}\cdot\delta{\bm{u}} =F(cos(θ)sin(θ)ρ˙(a)(1+ε)l˙(a))δa\displaystyle={F}\left(\cos(\theta)-\sin(\theta)\dot{\rho}({a})-(1+\varepsilon)\dot{l}({a})\right)\delta{a} (3.14)
12F(sin(θ)ρ¨(a)+(1+ε)l¨(a))(δa)2+o((δa)2).\displaystyle\qquad-\frac{1}{2}{F}\left(\sin(\theta)\ddot{\rho}({a})+(1+\varepsilon)\ddot{l}({a})\right)(\delta{a})^{2}+o\left((\delta{a})^{2}\right).

The symbol oo, that appears in (3.14) and elsewhere, is the Bachmann-Landau “Small-Oh” symbol. Its primary property of relevance is that o((δa)n)/(δa)n0o\left(\left(\delta a\right)^{n}\right)/(\delta a)^{n}\to 0, where n=0,1,n=0,~{}1,\ldots, as δa0\delta a\to 0.

Refer to caption
Figure 4: (a)–(c): Thin film peeling on a sinusoidal surface that does not involve contact. In this example, A=0.25A=0.25, λ=1.0\lambda=1.0, θ=π/3\theta=\pi/3, and ϱ(x1)=cos(2πx1)\varrho({x}_{1})=-\cos\left(2\pi{x}_{1}\right). (a) The peeling configurations corresponding to the equilibrium state as marked in (b). (b) The plot of the set of points 𝒟(F,ρ)×{F}\mathcal{D}^{\circ}(F,\rho)\times\{F\}. The peel-off force F+=0.099F^{+}=0.099, which is much greater than the peel-off force, 0.01, for peeling on a flat surface with the same nominal peeling angle. (c) The plot of the signed curvature and graph of ρ\rho of the sinusoidal surface. (d)–(f): Thin film peeling on a complicated surface that does not involve contact. In this example, A=1.0A=1.0, λ=1.0\lambda=1.0, θ=π/2\theta=\pi/2, ϱ(x1)=(sin(2πx1)+cos(4πx1+π/4)/2+sin(8πx1+π/3)+0.256)/1.315\varrho(x_{1})=\left(\sin(2\pi x_{1})+\cos(4\pi x_{1}+\pi/4)/2+\sin(8\pi x_{1}+\pi/3)+0.256\right)/1.315. (d) The peeling configurations corresponding to the equilibrium state as marked in (e). (e) The plot of the set of points 𝒟(F,ρ)×{F}\mathcal{D}^{\circ}(F,\rho)\times\{F\}. The peel-off force F+=0.97F^{+}=0.97, which is much greater than the peel-off force, 0.005, for peeling on a flat surface with the same nominal peeling angle. (f) The plot of the signed curvature and graph of ρ\rho of the complicated surface. In (b) and (e), the stable equilibrium state is marked as \bullet, neutral as \odot, and unstable as \circ.

In the current case of no contact it also follows that the true peeling angle is given by

ψ(a)=θ+tan1(ρ˙(a)).\psi(a)=\theta+\tan^{-1}\left(\dot{\rho}\left(a\right)\right). (3.15)

By recognizing from (3.15) that

cos(ψ(a))l˙(a;ρ)=cos(θ)sin(θ)ρ˙(a),\cos(\psi(a))\dot{l}({a};\rho)=\cos(\theta)-\sin(\theta)\dot{\rho}(a), (3.16)

we can rewrite (3.14) as

𝑭δ𝒖\displaystyle-{\bm{F}}\cdot\delta{\bm{u}} =F(cos(ψ(a))(1+ε))l˙(a;ρ)δa\displaystyle={F}\left(\cos(\psi(a))-(1+\varepsilon)\right)\dot{l}({a};\rho)\delta{a} (3.17)
12F(sin(θ)ρ¨(a)+(1+ε)l¨(a;ρ))(δa)2+o((δa)2).\displaystyle\qquad-\frac{1}{2}{F}\left(\sin(\theta)\ddot{\rho}({a})+(1+\varepsilon)\ddot{l}({a};\rho)\right)(\delta{a})^{2}+o((\delta{a})^{2}).

By substituting (3.17) into (3.8), and noting that on account of our non-dimensionalization scheme ε=F\varepsilon=F, we get that

δE(δa;a,F,ρ)\displaystyle\delta E(\delta{a};a,{F},\rho) =δE1(a;ρ,F)δa+δE2(a;ρ,F)(δa)2+o((δa)2),\displaystyle=\delta E_{1}(a\,;\rho,F)\delta a+\delta E_{2}(a\,;\rho,F)\left(\delta a\right)^{2}+o\left(\left(\delta a\right)^{2}\right), (3.18a)
where
δE1(a;ρ,F)\displaystyle\delta E_{1}(a\,;\rho,F) :=(12F2+F(cos(ψ(a))1)+w)l˙(a;ρ),\displaystyle:=\left(-\frac{1}{2}{F}^{2}+{F}(\cos(\psi(a))-1)+{w}\right)\dot{l}({a};\rho), (3.18b)
δE2(a;ρ,F)\displaystyle\delta E_{2}(a\,;\rho,F) :=12(Fsin(θ)ρ¨(a)+(F22+Fw)l¨(a;ρ)).\displaystyle:=-\frac{1}{2}\left({F}\sin(\theta)\ddot{\rho}({a})+\left(\frac{{F}^{2}}{2}+{F}-{w}\right)\ddot{l}({a};\rho)\right). (3.18c)

3.2.2 Equilibrium state

Since we take ρC2\rho\in C^{2}, it follows that ρ˙\dot{\rho} and l˙\dot{l} are continuous, and, as a consequence, that δE1(;ρ,F)\delta E_{1}(\cdot;\rho,F) is continuous. It then follows that for aa to be an equilibrium de-adhered length it is necessary that δE1(a;ρ,F)=0\delta E_{1}(a\,;\rho,F)=0. For a given FF and ρ\rho there can, however, be more than one de-adhered length that is in equilibrium. We characterize all those lengths by saying that they belong to the set

𝒟(F,ρ):={a𝒟|δE1(a;ρ,F)=0}.\mathcal{D}^{\circ}(F,\rho):=\left\{a\in\mathcal{D}~{}|~{}\delta E_{1}(a\,;\rho,F)=0\right\}. (3.19)

We plot the set of points

𝒟(F,ρ){F}\mathcal{D}^{\circ}(F,\rho)\mathbin{\leavevmode\hbox to6.46pt{\vbox to6.46pt{\pgfpicture\makeatletter\hbox{\hskip 3.22914pt\lower-3.22914pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}\pgfsys@setlinewidth{0.43056pt}\pgfsys@invoke{ }{}{{}}{} {}{}{}{{}}{} {}{}{}\pgfsys@moveto{-3.01387pt}{-3.01387pt}\pgfsys@lineto{3.01387pt}{3.01387pt}\pgfsys@moveto{-3.01387pt}{3.01387pt}\pgfsys@lineto{3.01387pt}{-3.01387pt}\pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}\{F\}

for various FF values, for the example surfaces shown in Figures 4a and c in Figures 4b and d, respectively. As can be seen, the peeling force for peeling on wavy surface is not constant, but varies with the same periodicity of the wavy surfaces. It can be shown that the point set 𝒟(F,ρ){F}\mathcal{D}^{\circ}(F,\rho)\mathbin{\leavevmode\hbox to6.46pt{\vbox to6.46pt{\pgfpicture\makeatletter\hbox{\hskip 3.22914pt\lower-3.22914pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}\pgfsys@setlinewidth{0.43056pt}\pgfsys@invoke{ }{}{{}}{} {}{}{}{{}}{} {}{}{}\pgfsys@moveto{-3.01387pt}{-3.01387pt}\pgfsys@lineto{3.01387pt}{3.01387pt}\pgfsys@moveto{-3.01387pt}{3.01387pt}\pgfsys@lineto{3.01387pt}{-3.01387pt}\pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}\{F\}, for any admissible FF, falls on the graph of the periodic function

F()\displaystyle F(\cdot) :=ψ,\displaystyle:=\mathpzc{F}\circ\psi, (3.20a)
where ψ\psi is defined in (3.15) and :[0,π]𝒟\mathpzc{F}:[0,\,\pi]\to\mathcal{D} is defined by the equation
(ψ)=cos(ψ)1\displaystyle\mathpzc{F}(\psi)=\cos\left(\psi\right)-1 +((cos(ψ)1)2+2w)1/2.\displaystyle+\left(\left(\cos(\psi)-1\right)^{2}+2{w}\right)^{1/2}. (3.20b)

We only consider cases in which the work of adhesion ww is non-negative. It therefore follows from (3.20b) that (ψ)\mathpzc{F}(\psi) is always non-negative. The graph of \mathcal{F} for different ww values is shown in Figure 5. As can be seen, \mathcal{F} is a strictly decreasing function whose value at any admissible ψ\psi increases with ww.

Kendall analyzed the peeling of a thin film on a flat smooth surface [19]. By setting ρ(x1)=0\rho(x_{1})=0 for a flat surface, which leads to ψ(a)=θ\psi(a)=\theta from (3.15), Kendall’s result can be immediately recovered from (3.20) which gives the peeling force as

F(a)=cos(θ)1+((cos(θ)1)2+2w)1/2,F(a)=\cos(\theta)-1+(\left(\cos(\theta)-1\right)^{2}+2{w})^{1/2}, (3.21)

which is a constant for a given nominal peeling angle θ\theta.

Refer to caption
Figure 5: The plot of the graph of \mathcal{F} as ψ\psi increases from 0 to π\pi for different w{w}.

For a general peeling process we define the supremum and infimum equilibrium force values, denoted as F+F^{+} and FF^{-}, respectively, as

F±=sup/inf{(ψ(𝒶))|𝒶𝒟}.F^{\pm}=\sup/\inf\,\{\mathpzc{F}\left(\psi(a)\right)~{}|~{}a\in\mathcal{D}\}. (3.22)

It can be shown that the maximum and minimum values of the function \mathpzc{F} are (2w)1/2(2w)^{1/2} and (4+2w)1/22(4+2w)^{1/2}-2, respectively. This implies that F+F^{+} is bounded above by (2w)1/2(2w)^{1/2} and FF^{-} is bounded below by (4+2w)1/22(4+2w)^{1/2}-2.

We denote the maximum and minimum values of the true peeling angle during peeling as ψ+\psi^{+} and ψ\psi^{-}, respectively. In the current case of peeling with no contact it follows from the fact that \mathcal{F} is a monotonically decreasing function that

F±\displaystyle F^{\pm} =(ψ),\displaystyle=\mathpzc{F}\left(\psi^{\mp}\right), (3.23a)
where
ψ±:=θ\displaystyle\psi^{\pm}:=\theta +tan1(ρ˙±).\displaystyle+\tan^{-1}\left(\dot{\rho}^{\pm}\right). (3.23b)

Since (ψ)\mathpzc{F}(\psi) is always non-negative it follows from (3.23a) that F±0F^{\pm}\geq 0.

Theorem 3.1.

If F[F,F+]F\notin[F^{-},\;F^{+}] then there does not exist any equilibrium de-adhered length at that FF (e.g., see Figures 4b and e).

Proof.

Since ψ±\psi^{\pm} are the maximum and minimum values of ψ(a)\psi(a), respectively, and from (3.9) we have

0ψψ(a)ψ+π.0\leq\psi^{-}\leq\psi(a)\leq\psi^{+}\leq\pi. (3.24)

From (3.24) and the fact that the cosine function is strictly decreasing in the interval [0,π][0,\pi], we find that

1cos(ψ+)cos(ψ(a))cos(ψ)1.-1\leq\cos\left(\psi^{+}\right)\leq\cos\left(\psi(a)\right)\leq\cos\left(\psi^{-}\right)\leq 1. (3.25)

The inequalities (3.25) allow us to express cos(ψ(a))\cos\left(\psi(a)\right) as either cos(ψ)δ\cos\left(\psi^{-}\right)-\delta^{-} or cos(ψ+)+δ+\cos\left(\psi^{+}\right)+\delta^{+}; and, though they depend on aa, δ\delta^{\mp}, are always non-negative.

Let us first consider the case F>F+F>F^{+}. When FF is strictly greater than F+F^{+} it can be represented as F++ϵ+F^{+}+\epsilon^{+} for some ϵ+>0\epsilon^{+}>0. Using this representation for FF and expressing cos(ψ(a))\cos\left(\psi(a)\right) as cos(ψ)δ\cos\left(\psi^{-}\right)-\delta^{-} in (3.18b) we get that

δE1(a;ρ,F)=12(2F+(δ+ϵ+)+ϵ+(2δ+ϵ++2))l˙(a;ρ)+cos(ψ)ϵ+l˙(a;ρ).\delta E_{1}(a\,;\rho,F)=-\frac{1}{2}\left(2F^{+}\left(\delta^{-}+\epsilon^{+}\right)+\epsilon^{+}\left(2\delta^{-}+\epsilon^{+}+2\right)\right)\dot{l}(a;\rho)+\cos\left(\psi^{-}\right)\epsilon^{+}\ \dot{l}(a;\rho). (3.26)

Equation (3.4) implies that the derivative of the peeled length l˙(a;ρ)\dot{l}(a;\rho) is always positive. For that reason and since cos(ψ)1\cos\left(\psi^{-}\right)\geq-1 it follows from (3.26) that when F>F+F>F^{+}

δE1(a;ρ,F)12(2F+(δ+ϵ+)+ϵ+(2δ+ϵ+))l˙(a;ρ).\delta E_{1}(a\,;\rho,F)\leq-\frac{1}{2}\left(2F^{+}\left(\delta^{-}+\epsilon^{+}\right)+\epsilon^{+}\left(2\delta^{-}+\epsilon^{+}\right)\right)\dot{l}(a;\rho). (3.27)

It follows from (3.27) and the facts that l˙(a;ρ)\dot{l}(a;\rho) and ϵ+\epsilon^{+} are positive and δ\delta^{-} and F+F^{+} are non-negative that when F>F+F>F^{+}

δE1(a;ρ,F)<0.\delta E_{1}(a;\rho,F)<0. (3.28)

Since for equilibrium it is necessary that δE1(a;ρ,F)=0\delta E_{1}(a;\rho,F)=0 it follows from (3.28) that there can exist no equilibrium de-adhered lengths when F>F+F>F^{+}.

Now consider the case 0F<F0\leq F<F^{-}. When F<FF<F^{-} it can be represented as FϵF^{-}-\epsilon^{-} for some ϵ>0\epsilon^{-}>0. Representing FF this way and expressing cos(ψ(a))\cos\left(\psi(a)\right) as cos(ψ+)+δ+\cos\left(\psi^{+}\right)+\delta^{+} in (3.18b) we determine that

δE1(a;ρ,F)=12(2F(δ++ϵ)ϵ(2δ++ϵ2))l˙(a;ρ)ϵcos(ψ+)l˙(a;ρ).\delta E_{1}(a;\rho,F)=\frac{1}{2}\left(2F^{-}\left(\delta^{+}+\epsilon^{-}\right)-\epsilon^{-}\left(2\delta^{+}+\epsilon^{-}-2\right)\right)\dot{l}(a;\rho)-\epsilon^{-}\cos\left(\psi^{+}\right)\dot{l}(a;\rho). (3.29)

Since l˙(a;ρ)>0\dot{l}(a;\rho)>0 and cos(ψ+)1\cos\left(\psi^{+}\right)\leq 1 it follows from (3.29) that when F<FF<F^{-}

δE1(a;ρ,F)12(2F(δ++ϵ)ϵ(2δ++ϵ))l˙(a;ρ),\delta E_{1}(a;\rho,F)\geq\frac{1}{2}\left(2F^{-}\left(\delta^{+}+\epsilon^{-}\right)-\epsilon^{-}\left(2\delta^{+}+\epsilon^{-}\right)\right)\dot{l}(a;\rho),

which can be re-arranged to read

δE1(a;ρ,F)12(2F(δ++ϵ)+(ϵ)2)l˙(a;ρ).\delta E_{1}(a;\rho,F)\geq\frac{1}{2}\left(2F\left(\delta^{+}+\epsilon^{-}\right)+\left(\epsilon^{-}\right)^{2}\right)\dot{l}(a;\rho). (3.30)

Recalling that δ+\delta^{+} is non-negative and ϵ\epsilon^{-} is positive it follows from (3.30) that when 0F<F0\leq F<F^{-}

δE1(a;ρ,F)>0.\delta E_{1}(a;\rho,F)>0. (3.31)

For the same reason as before, the inequality (3.31) implies that when FF is less than FF^{-} but still non-negative then there cannot exist any de-adhered lengths that are in equilibrium. ∎

The inequality (3.28) implies that the derivative of the total energy w.r.t aa will be negative when FF is greater than F+F^{+} irrespective of the value of aa, ww, or the nature of ρ\rho (see, e.g., Figures 6a and c). Thus, if F>F+F>F^{+} the de-adhered length will grow without bound. Realistically, however, the de-adhered length will keep growing until the film completely detaches from the substrate. For that reason we call F+F^{+} the peel-off force.

The result that (3.31) holds in the case where 0F<F0\leq F<F^{-} implies that the derivative of the total energy with respect to the de-adhered length in that case is positive irrespective of any other details in the problem (see, for example, Figures 6b and d). Therefore, if 0F<F0\leq F<F^{-} and the de-adhered length is initially naught, then the de-adhered length will not grow or if it is initially non-zero then it will keep decreasing, i.e., the peel front will keep receding until the entire thin film is adhered to the substrate. For this reason, we call FF^{-} the peel-initiation force.

Refer to caption
Figure 6: The plot of potential energy as a function of a{a} for different values of F{F} for peeling on the sinusoidal [(a)–(b)] and complicated [(c)–(d)] surfaces. Note that there is no local minima of the energy when F>F+F>F^{+} in (a) and (c). The labels ①–③ correspond to those marked in Figures 4b and d. There is also no local minima of the energy when 0F<F0\leq F<F^{-} in (b) and (d).

3.2.3 Stability of equilibrium state

We study the stability of the local equilibria by examining the sign of the second variation of the total potential energy. Specifically, a configuration with the de-adhered length aa is a stable equilibrium state iff aa belongs to the set

𝒟(F,ρ):={a𝒟(F,ρ)|δE2(a;ρ,F)>0}.\mathcal{D}^{\leavevmode\hbox to1.9pt{\vbox to1.9pt{\pgfpicture\makeatletter\hbox{\hskip 0.95pt\lower-0.95pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{}{{{}} {}{}{}{}{}{}{}{} }\pgfsys@beginscope\pgfsys@invoke{ }\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{0.75pt}{0.0pt}\pgfsys@curveto{0.75pt}{0.41422pt}{0.41422pt}{0.75pt}{0.0pt}{0.75pt}\pgfsys@curveto{-0.41422pt}{0.75pt}{-0.75pt}{0.41422pt}{-0.75pt}{0.0pt}\pgfsys@curveto{-0.75pt}{-0.41422pt}{-0.41422pt}{-0.75pt}{0.0pt}{-0.75pt}\pgfsys@curveto{0.41422pt}{-0.75pt}{0.75pt}{-0.41422pt}{0.75pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}(F,\rho):=\left\{a\in\mathcal{D}^{\circ}(F,\rho)~{}|~{}\delta E_{2}(a\,;\rho,F)>0\right\}. (3.32)

It is a neutral equilibrium state iff aa belongs to the set

𝒟(F,ρ):={a𝒟(F,ρ)|δE2(a;ρ,F)=0},\mathcal{D}^{\odot}(F,\rho):=\left\{a\in\mathcal{D}^{\circ}(F,\rho)~{}|~{}\delta E_{2}(a\,;\rho,F)=0\right\}, (3.33)

and is an unstable equilibrium state if aa belongs to the set 𝒟(F,ρ)\mathcal{D}^{\otimes}(F,\rho), which is a set of all de-adhered lengths that belong to 𝒟(F,ρ)\mathcal{D}^{\circ}(F,\rho) but not to 𝒟(F,ρ)\mathcal{D}^{\leavevmode\hbox to1.9pt{\vbox to1.9pt{\pgfpicture\makeatletter\hbox{\hskip 0.95pt\lower-0.95pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{}{{{}} {}{}{}{}{}{}{}{} }\pgfsys@beginscope\pgfsys@invoke{ }\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{0.75pt}{0.0pt}\pgfsys@curveto{0.75pt}{0.41422pt}{0.41422pt}{0.75pt}{0.0pt}{0.75pt}\pgfsys@curveto{-0.41422pt}{0.75pt}{-0.75pt}{0.41422pt}{-0.75pt}{0.0pt}\pgfsys@curveto{-0.75pt}{-0.41422pt}{-0.41422pt}{-0.75pt}{0.0pt}{-0.75pt}\pgfsys@curveto{0.41422pt}{-0.75pt}{0.75pt}{-0.41422pt}{0.75pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}(F,\rho) or 𝒟(F,ρ)\mathcal{D}^{\odot}(F,\rho).

Stability and surface curvature

Suppose aa is an equilibrium de-adhered length at the force FF. Then it is necessary that FF and aa satisfy the equation F=F(a)F=F(a), where F()F(\cdot) on the right hand side is the function defined in (3.20). Upon substituting FF in (3.18c) with F(a)F(a), and then simplifying the resulting equation, we get that

δE2(a;ρ,F(a))=F(a)2k(a)l˙(a;ρ)2sin(ψ(a)),\delta E_{2}(a\,;\rho,F(a))=-\frac{F(a)}{2}k({a})\dot{l}(a;\rho)^{2}\sin\left(\psi(a)\right), (3.34)

where

k(a)=ρ¨(a)/l˙(a;ρ)3k({a})=\ddot{\rho}({a})/\dot{l}(a;\rho)^{3} (3.35)

is the signed curvature of the graph of ρ\rho. The mean curvature of the substrate’s surface at the point whose coordinates w.r.t to \mathbscre1\mathbscr{e}_{1} and \mathbscre2\mathbscr{e}_{2} are aa and ρ(a)\rho(a), respectively, equals k(a)/(2λ)k(a)/(2\lambda). Therefore, we will often refer to k(a)k(a) as the substrate’s surface curvature.

If w=0w=0, then from (3.20) we have that F(a)=0F(a)=0. Consequently, from (3.34), δE2=0\delta E_{2}=0 for all aa. Therefore, when w=0w=0 all states are neutral-equilibrium states. In the following section we take that ww is positive (recall that w0w\geq 0).

If k(a)k(a) vanishes then it follows from (3.34) that aa belongs to 𝒟\mathcal{D}^{\odot}, i.e., that the corresponding state is a neutral equilibrium state.

It follows from the definitions of a±a^{\pm}, ρ˙±\dot{\rho}^{\pm}, and kk and the smoothness of ρ\rho that k(a)=0k\left(a\right)=0 iff a=a±a=a^{\pm}. So, if k(a)k(a) does not vanish then aa is different from a±a^{\pm}, which implies from the definitions of ρ˙±\dot{\rho}^{\pm}, ψ\psi, and ψ±\psi^{\pm} that ψ(a)\psi(a) is different from ψ±\psi^{\pm}. This last deduction in conjunction with (3.24) implies that when k(a)k(a) is not naught, sin(ψ(a))\sin(\psi(a)) is positive. Hence, it follows from (3.34) that the configuration is stable (resp. unstable) when k(a)k(a) is negative (resp. positive). These results, which connect an equilibrium state’s stability to the surface curvature, are illustrated by Figures 4b–c for a sinusoidal surface and by Figures 4e–f for a complicated surface.

Positive, negative, and zero values of the function F˙()\dot{F}(\cdot) imply stable, unstable, and neutral equilibria, respectively

Again, let aa be an equilibrium de-adhered length at the force FF. From (3.20) we have that

F˙(a)=k(a)l˙(a;ρ)sin(ψ(a))(11cos(ψ(a))(2w+(1cos(ψ(a)))2)1/2).\dot{F}(a)=-k(a)\dot{l}(a;\rho)\sin(\psi(a))\left(1-\frac{1-\cos(\psi(a))}{(2{w}+(1-\cos(\psi(a)))^{2})^{1/2}}\right). (3.36)

Recall that we deduced that when w=0w=0 all configurations are neutral equilibrium configurations. Therefore, in the following two paragraphs we take that w>0w>0.

Say F˙(a)\dot{F}(a) vanishes. It can be checked using (3.13a) that l˙(a;ρ)\dot{l}(a;\rho) is always positive, and since we have assumed that w>0w>0 it can be shown that the expression within the large parenthesis on the right hand side of (3.36) is always positive. Therefore, if F˙(a)\dot{F}(a) vanishes then we have the following three cases from (3.36): (i) the factor k(a)k(a) vanishes, (ii) the factor sin(ψ(a))\sin(\psi(a)) vanishes, (iii) both these factors vanish. The factor k(a)k(a) vanishes in both case (i) and (iii). Let us focus on case (ii). If sin(ψ(a))\sin(\psi(a)) vanishes then we know from (3.24) that a=a±a=a^{\pm}, which then implies, based on the discussion following (3.34), that k(a)k(a) has to also vanish. then we know from (3.24) that a=a±a=a^{\pm}, which then implies, based on the discussion contained in the third paragraph following the one containing (3.34) that, k(a)k(a) has to also vanish. Thus, k(a)k(a) vanishes in all three cases. That is, if F˙(a)\dot{F}(a) is equal to zero then k(a)k(a) is also equal to zero. This last deduction in light of the results presented in Stability and surface curvature implies that if F˙(a)\dot{F}(a) vanishes then the configuration corresponding to aa is a neutral-equilibrium configuration.

Now say that F˙(a)\dot{F}(a) is positive (resp. negative). As previously stated, since we have assumed that w>0w>0 the expression within the large paranthesis on the right hand side of (3.36) is always positive. The factor sin(ψ(a))\sin(\psi(a)) is positive since we can show using (3.24) that it is always non-negative and if it were to vanish then that would contradict the assumption that F˙(a)\dot{F}(a) is non-zero. This, in conjuction with (3.36), imply that k(a)k(a) is negative (resp. positive) whenever F˙(a)\dot{F}(a) is positive (resp. negative). In light of the results presented in Stability and surface curvature, this last deduction implies that when F˙(a)\dot{F}(a) is positive (resp. negative) then the corresponding equilibrium configuration is stable (resp. unstable).

These results are illustrated in Figures 4b and e using the example surface profiles shown in Figures 4c and f.

3.3 Peeling process that might involve contact

Refer to caption
Figure 7: A schematic of peeling with contact. (a) Forward peeling, non-local contact. (b) Backward peeling, non-local contact. (c) Forward peeling, local contact. (d) Backward peeling, local contact.

If θ\theta is kept fixed during the peeling process and that constant θ\theta violates the global compatibility condition (3.10), then there will exist some configurations during the peeling process that will involve contact777 This, of course, does not mean that in such a peeling process all configurations will involve contact. That is, there can exist configurations that involve no contact during parts of the peeling process (see, e.g., subfigure (b) in Figure 8). . We provide a procedure for determining whether or not a given configuration involves contact in §3.3.2. If the configuration 𝜿\bm{\kappa} does not involve contact, the results needed to compute the force FF so that (𝜿,F)(\bm{\kappa},F) is an equilibrium state and the results needed to determine the stability of (𝜿,F)(\bm{\kappa},F) are given in §3.2.

When 𝜿\bm{\kappa} involves contact the primitive conditions that determines whether or not a state (𝜿,F)(\bm{\kappa},F) is in equilibrium remain the same as before. Specifically, even when a configuration 𝜿\bm{\kappa} involves contact, the state (𝜿,F)(\bm{\kappa},F) is an equilibrium state iff the de-adhered length aa in 𝜿\bm{\kappa} belongs to the set 𝒟(F,ρ)\mathcal{D}^{\circ}(F,\rho), which is defined in (3.19). Similarly, it qualifies as a stable, neutral, or unstable state depending on whether the de-adhered length aa belongs to the sets 𝒟(F,ρ)\mathcal{D}^{\bullet}(F,\rho), 𝒟(F,ρ)\mathcal{D}^{\odot}(F,\rho), or 𝒟(F,ρ)\mathcal{D}^{\otimes}(F,\rho), respectively. These sets are defined in §3.2.3.

What makes the analysis of the case involving contact more challenging is the calculation of the functions δE1\delta E_{1} and δE2\delta E_{2}, which are needed for the construction of the sets 𝒟(F,ρ)\mathcal{D}^{\circ}(F,\rho), 𝒟(F,ρ)\mathcal{D}^{\bullet}(F,\rho), etc. These functions are defined in (3.18a) in terms of the asymptotic expansion of δE\delta E as δa0\delta a\to 0. The calculation of δE\delta E’s asymptotic expansion is challenging due to the presence of the term 𝑭δ𝒖-\bm{F}\cdot\delta\bm{u} in (3.8). In the case involving contact the term 𝑭δ𝒖-\bm{F}\cdot\delta\bm{u} simplifies to Fδu-F\delta u, where δu\delta u is δ𝒖\delta\bm{u}’s magnitude. This is due to the fact that δ𝒖\delta\bm{u} is always in the same direction as 𝑭\bm{F} (see Figures 12 and 13) when a configuration involves contact. This remains true irrespective of whether the contact region consists of a single contact patch (e.g., see Figure 7) or several contact patches (e.g., see subfigure (a) in either Figure 12 or 13). Thus, calculation of δE\delta E’s asymptotic expansion requires the calculations of δu\delta u’s asymptotic expansion, which in this current case is non-trivial. To elaborate, in the case not involving contact, it is straightforward to determine the asymptotic expansion of δu\delta u, e.g. through the use of (3.11) and (3.12). While now in the case involving contact this exercise is relatively more difficult.

We could not obtain a general, closed-form expression for the asymptotic expansion of δu\delta u when the configuration involved contact. However, in §3.3.3 we present a family of four analytical, but not closed-form, expressions for calculating δu\delta u’s asymptotic expansion, and from that δE1\delta E_{1} and δE2\delta E_{2}, that apply to special categories of contact configurations. We describe these four categories, to whom we henceforth refer to as C.1, C.2, etc., shortly in §3.3.1, but we note here that it will follow from their definitions that any contact configuration can be uniquely placed into in one of them.

From the family of δE1\delta E_{1} functions given in §3.3.3, which apply to different categories of contact configurations, we found that, interestingly, irrespective of which category a contact configuration 𝜿\bm{\kappa} belongs to the force FF needed to make the state (𝜿,F)(\bm{\kappa},F) an equilibrium state is always (ψ(a))\mathcal{F}(\psi(a)), where \mathcal{F} is defined in (3.20b). However, it follows from the family of δE2\delta E_{2} functions given in §3.3.3 that 𝜿\bm{\kappa}’s category is still relevant for determining the nature of (𝜿,F(a))(\bm{\kappa},F(a))’s stability.

In summary, our method for simulating a peeling process in which θ\theta violates the global compatibility condition at some stage of the peeling process is as follows. Let the peeling experiment be defined by prescribing the sequence (ai,θi)(a_{i},\theta_{i}), where i=1i=1, 2,2, etc., and the symbols aia_{i} and θi\theta_{i} are the de-adhered length and the nominal peel angle, respectively, in the ithi^{\rm th} step of the experiment. We compute ψ(ai)\psi(a_{i}), the true peeling angle for the ithi^{\rm th} step, using Algorithm 1. We place the configuration 𝜿i\bm{\kappa}_{i} into one of the four categories, C.1–4, using θi\theta_{i} and ψ(ai)\psi(a_{i}) (see §3.3.2 for details). We then compute the force FiF_{i} such that the state 𝒮i=(𝜿i,Fi)\mathcal{S}_{i}=(\bm{\kappa}_{i},F_{i}) becomes an equilibrium state as (ψ(ai))\mathcal{F}(\psi(a_{i})). We determine the nature of 𝒮i\mathcal{S}_{i}’s stability by computing δE2(ai;ρ,Fi)\delta E_{2}(a_{i};\rho,F_{i}) and constructing the sets 𝒟\mathcal{D}^{\leavevmode\hbox to1.9pt{\vbox to1.9pt{\pgfpicture\makeatletter\hbox{\hskip 0.95pt\lower-0.95pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{}{{{}} {}{}{}{}{}{}{}{} }\pgfsys@beginscope\pgfsys@invoke{ }\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@moveto{0.75pt}{0.0pt}\pgfsys@curveto{0.75pt}{0.41422pt}{0.41422pt}{0.75pt}{0.0pt}{0.75pt}\pgfsys@curveto{-0.41422pt}{0.75pt}{-0.75pt}{0.41422pt}{-0.75pt}{0.0pt}\pgfsys@curveto{-0.75pt}{-0.41422pt}{-0.41422pt}{-0.75pt}{0.0pt}{-0.75pt}\pgfsys@curveto{0.41422pt}{-0.75pt}{0.75pt}{-0.41422pt}{0.75pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}, 𝒟\mathcal{D}^{\odot}, etc. When the contact configuration 𝜿\bm{\kappa} belongs to either C.1 or C.2 the Algorithm 1 also provides the value of the parameter \ell, which is the distance between the peeling and contact fronts. So, when 𝜿\bm{\kappa} belongs to either C.1 or C.2 we use that value in conjunction with (3.40b) to compute δE2(ai;ρ,Fi)\delta E_{2}(a_{i};\rho,F_{i}). When 𝜿\bm{\kappa} belongs to C.3 or C.4 we compute the value of δE2\delta E_{2} using (3.42b) or (3.44b), respectively.

We demonstrate our method by using the same two example surfaces that we previously considered in §3.2. The schematics of these simple and complicated wavy surfaces are shown, e.g., in Figures 4a and d, respectively.

On each surface we simulated two (virtual) peeling experiments. In the first experiment—forward peeling (defined in §3.3.1)—the peeling angle was kept fixed at a value of π/3\pi/3 through out the experiment, while in the second one—backward peeling (defined in §3.3.1)—it was kept fixed at 3π/43\pi/4. The constant nominal peeling angles we chose, namely π/3\pi/3 and 3π/43\pi/4, violated the global compatibility condition on both our example surfaces. Therefore, the results from §3.2 cannot be used to simulate these experiments, for instance to generate the set

F[F,F+]𝒟(F,ρ){F}\bigcup_{F\in[F^{-},F^{+}]}\mathcal{D}^{\circ}(F,\rho)\mathbin{\leavevmode\hbox to6.46pt{\vbox to6.46pt{\pgfpicture\makeatletter\hbox{\hskip 3.22914pt\lower-3.22914pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}\pgfsys@setlinewidth{0.43056pt}\pgfsys@invoke{ }{}{{}}{} {}{}{}{{}}{} {}{}{}\pgfsys@moveto{-3.01387pt}{-3.01387pt}\pgfsys@lineto{3.01387pt}{3.01387pt}\pgfsys@moveto{-3.01387pt}{3.01387pt}\pgfsys@lineto{3.01387pt}{-3.01387pt}\pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}\{F\} (3.37)

for these experiments. Note that a typical point in (3.37) represents an equilibrium state, with its abscissa denoting the state’s de-adhered length and its ordinate the state’s force. Therefore, we applied our method, which we introduced earlier in this section, to the sequence (ai,θ0)(a_{i},\theta_{0}) (θ0=π/3,3π/4\theta_{0}=\pi/3,~{}3\pi/4) and computed the sequence (Fi)(F_{i}), where FiF_{i} is the equilibrium force corresponding to aia_{i}; And constructed (a subset of) (3.37), alternately, as {(ai,Fi)}\left\{(a_{i},F_{i})\right\}. The sets (3.37) that we generated this way for the forward and backward peeling cases are shown in subfigures (b) and (e), respectively, of Figure 8 for the simple wavy surface and in Figure 9 for the complicated wavy surface.

Note that our method also determines the stability of an equilibrium state and informs us whether or not that state involves contact. In the subfigures (b) and (e) of Figures 8 and 9 the stable equilibrium states are denoted using solid/filled symbols while unstable states are denoted using hollow/unfilled symbols. In the subfigures we identify the states that involve contact by placing them over a yellow background. As can be noted from the subfigure, a yellow region is preceded and followed by white regions, and vice versa. Thus, when the global compatibility condition is violated a sequence of configurations involving contact can be followed by a sequence of configurations not involving contact, and so on.

Finally, our method also provides the true peeling angle sequences (ψ(ai))\left(\psi(a_{i})\right) in the experiments. These are shown in subfigures (a) and (b) of Figures 8 and 9.

A few representative configurations from the peeling experiments are explicitly sketched in subfigures (c) and (f) of Figures 8 and 9.

3.3.1 The four categories of contact configurations

A configuration involving contact can be placed into one of the following four categories.

  1. C.1

    Forward-peeling, non-local contact (Figure 7a),

  2. C.2

    Backward-peeling, non-local contact (Figure 7b),

  3. C.3

    Forward-peeling, local contact (Figure 7c), and

  4. C.4

    Backward-peeling, local contact (Figure 7d).

We call a configuration a forward-peeling configuration if the θ\theta in it is less than π/2\pi/2, and a backward-peeling configuration otherwise. Roughly speaking, we consider configurations of the type shown in Figures 7a and b as those that involve non-local contact, and configurations of the type shown in Figures 7c and d as those that involve local contact. We define local and non-local contact precisely by introducing the notions of contact region and contact front, which we discuss next.

We define the contact region corresponding to the deformed configuration 𝜿\bm{\kappa} as Γc={x1𝒟|𝒙(x1)𝕊}\Gamma_{c}=\{x_{1}\in\mathcal{D}~{}|~{}\bm{x}(x_{1})\in\partial\mathbb{S}\}, where

𝕊={x1\mathbscre1+x2\mathbscre2+x3\mathbscre3𝔼|(x1,x2,x3)3andx2=ρ(x1)}\partial\mathbb{S}=\left\{x_{1}\mathbscr{e}_{1}+x_{2}\mathbscr{e}_{2}+x_{3}\mathbscr{e}_{3}\in\mathbb{E}~{}|~{}(x_{1},x_{2},x_{3})\in\mathbb{R}^{3}~{}\text{and}~{}x_{2}=\rho(x_{1})\right\} (3.38)

is the substrate’s surface (cf. (3.1)) and 𝒙\bm{x} is defined in (3.3b). Let cc be the point in Γc\Gamma_{c} that is closest to Γa\Gamma_{a} in 𝒟\mathcal{D}’s topology; recall here that Γa\Gamma_{a} is the adhered region corresponding to the configuration 𝜿\bm{\kappa}. We define the contact front as C={O+𝒙(c)+x3\mathbscre3||x3|b/2}C=\{O+\bm{x}(c)+x_{3}\mathbscr{e}_{3}\in\mathcal{E}~{}|~{}\lvert x_{3}\rvert\leq b/2\}.

We say that a deformed configuration 𝜿\bm{\kappa} involves local contact iff C=PC=P and involves non-local contact otherwise.

3.3.2 Procedure for determining a configuration’s type—contact or non-contact—and a contact configuration’s category

Algorithm 1 Procedure for determining the category of a configuration
1:procedure Generate (true peeling angle ψ(a)\psi(a) and distance \ell)
2:     Input: Substrate’s surface profile ρ\rho, nominal peeling angle θ\theta, and de-adhered length aa
3:     Let Δρ(x1):=ρ(a)tan(θ)(x1a)ρ(x1)\Delta\rho(x_{1}):=\rho(a)-\tan\left(\theta\right)\left(x_{1}-a\right)-\rho\left(x_{1}\right), where x1x_{1}\in\mathbb{R}.
4:     if Δρ(x1)>0\Delta\rho(x_{1})>0 for all x1x_{1} satisfying sgn((θπ/2)(x1a))>0\text{sgn}\left((\theta-\pi/2)(x_{1}-a)\right)>0888We define the sgn()\text{sgn}(\cdot) function as: sgn(x)={+1,x0,1,x<0.\text{sgn}(x)=\begin{cases}+1,&x\geq 0,\\ -1,&x<0.\end{cases}  then \triangleright No contact
5:         
ψ(a)=θ+tan1(ρ˙(a)).\psi(a)=\theta+\tan^{-1}(\dot{\rho}(a)).
\triangleright c.f. (3.15)
6:     else
7:         if k(a)0k(a)\leq 0 and ρ˙(a)sgn(θπ/2)>0\dot{\rho}(a)\text{sgn}(\theta-\pi/2)>0 then \triangleright Local type contact
8:              
ψ(a)={0,θ<π/2,π,θπ/2.\psi(a)=\begin{cases}0,&\theta<\pi/2,\\ \pi,&\theta\geq\pi/2.\end{cases}
9:         else\triangleright Non-local type contact
10:              Find the abscissa of contact front cc such that
ρ(a)ρ(c)ac=ρ˙(c)forsgn((θπ/2)(ca))>0.\frac{\rho(a)-\rho(c)}{a-c}=\dot{\rho}(c)\quad\text{for}\quad\mathrm{sgn}\left((\theta-\pi/2)(c-a)\right)>0.
11:              Then
ψ(a)={tan1(ρ˙(a))tan1(ρ˙(c)),θ<π/2, forward peeling,π+tan1(ρ˙(a))tan1(ρ˙(c)),θπ/2, backward peeling,\psi(a)=\begin{cases}\tan^{-1}\left(\dot{\rho}(a)\right)-\tan^{-1}\left(\dot{\rho}(c)\right),&\theta<\pi/2,\text{ forward peeling,}\\ \pi+\tan^{-1}\left(\dot{\rho}(a)\right)-\tan^{-1}\left(\dot{\rho}(c)\right),&\theta\geq\pi/2,\text{ backward peeling,}\end{cases}
12:              and
=((ac)2+(ρ(a)ρ(c))2)1/2.\ell=\left((a-c)^{2}+\left(\rho(a)-\rho(c)\right)^{2}\right)^{1/2}.
13:         end if
14:     end if
15:end procedure

When θ\theta satisfies the global compatibility condition then we know that the configuration 𝜿\bm{\kappa} will be of non-contact type. When the global compatibility condition is violated, as we describe in the next few paragraphs, determining whether or not 𝜿\bm{\kappa} involves contact, i.e., determining its type, and if it involves contact then determining the contact category that 𝜿\bm{\kappa} belongs to essentially comes down to determining the true peeling angle ψ(a)\psi(a).

The true peeling angle is defined in §3.1.3. When θ\theta satisfies the global compatibility condition ψ(a)\psi(a) is given by (3.15). When θ\theta violates the the global compatibility condition it can be computed using the numerical procedure that we present in Algorithm 1.

Given a configuration 𝜿\bm{\kappa}, if the true peeling angle ψ(a)\psi(a) in it is different from  (3.15) then 𝜿\bm{\kappa} is a contact type configuration. Otherwise, it is a non-contact type configuration.

For placing a contact type configuration 𝜿\bm{\kappa} into one of the four categories described in §3.3.1 it is sufficient to know whether 𝜿\bm{\kappa} is a forward or backward peeling configuration and whether the contact in it is of the local or the non-local type.

The configuration 𝜿\bm{\kappa} is a forward peeling configuration if θ<π/2\theta<\pi/2, and a backward peeling configuration otherwise. A contact configuration 𝜿\bm{\kappa} involves non-local or local contact depending on whether the true peeling angle in it, ψ(a)\psi(a), lies in the interior or on the boundary of the set [0,π][0,\pi].

3.3.3 Asymptotic expansion of δu\delta u and the functions δE1\delta E_{1} and δE2\delta E_{2} for the different categories of contact configurations

Categories C.1–2

In §A we show that for these categories

δu=\displaystyle\delta{u}= (1+εcos(ψ(a)))l˙(a;ρ)δa\displaystyle\left(1+\varepsilon-\cos(\psi(a))\right)\dot{l}({a};\rho)\delta{a} (3.39)
+12((1+ε)l¨(a;ρ)l˙(a;ρ)2sin2(ψ(a))+(sin(ψ(a))cos(ψ(a))ρ˙(a))ρ¨(a)l˙(a;ρ))(δa)2+o((δa)2),\displaystyle+\frac{1}{2}\left((1+\varepsilon)\ddot{l}({a};\rho)-\frac{\dot{l}({a};\rho)^{2}}{\ell}\sin^{2}(\psi(a))+\frac{\left(\sin(\psi(a))-\cos(\psi(a))\dot{\rho}(a)\right)\ddot{\rho}(a)}{\dot{l}(a;\rho)}\right)(\delta{a})^{2}+o((\delta{a})^{2}),

where \ell is the distance between PP and CC. Recall that PP and CC denote the peeling and contact front, respectively (e.g., see Figure 7). We introduced these notions in §3.1.1 and §3.3.1. Substituting the δu\delta u appearing in (3.8) with the expression appearing on the right hand side of (3.39) and then comparing the resulting equation with (3.18a) we get that

δE1(a;ρ,F)\displaystyle\delta E_{1}(a\,;\rho,F) :=(12F2+F(cos(ψ(a))1)+w)l˙(a;ρ),\displaystyle:=\left(-\frac{1}{2}{F}^{2}+{F}(\cos(\psi(a))-1)+{w}\right)\dot{l}({a};\rho), (3.40a)
δE2(a;ρ,F)\displaystyle\delta E_{2}(a\,;\rho,F) :=12((F22+Fw)l¨(a;ρ)F(l˙(a;ρ)2sin2(ψ(a))(sin(ψ(a))cos(ψ(a))ρ˙(a))ρ¨(a)l˙(a;ρ))).\displaystyle:=-\frac{1}{2}\left(\left(\frac{{F}^{2}}{2}+{F}-{w}\right)\ddot{l}({a};\rho)-F\left(\frac{\dot{l}(a;\rho)^{2}}{\ell}\sin^{2}(\psi(a))-\frac{\left(\sin(\psi(a))-\cos(\psi(a))\dot{\rho}(a)\right)\ddot{\rho}(a)}{\dot{l}(a;\rho)}\right)\right). (3.40b)
Categories C.3

As can be noted from Figure 12c for this category, δu=εδl\delta u=\varepsilon\delta l. With (3.12) we have

δu=ε(l˙(a;ρ)δa+12l¨(a;ρ)(δa)2+o((δa)2)).\delta{u}=\varepsilon\left(\dot{l}({a};\rho)\delta{a}+\frac{1}{2}\ddot{l}({a};\rho)(\delta{a})^{2}+o\left((\delta{a})^{2}\right)\right). (3.41)

As before, substituting δu\delta u from (3.41) into (3.8) and comparing the resulting equation with (3.18a) we get that

δE1(a;ρ,F)\displaystyle\delta E_{1}(a\,;\rho,F) :=(12F2w)l˙(a;ρ),\displaystyle:=-\left(\frac{1}{2}{F}^{2}-{w}\right)\dot{l}({a};\rho), (3.42a)
δE2(a;ρ,F)\displaystyle\delta E_{2}(a\,;\rho,F) :=12(F22w)l¨(a;ρ).\displaystyle:=-\frac{1}{2}\left(\frac{{F}^{2}}{2}-{w}\right)\ddot{l}({a};\rho). (3.42b)
Categories C.4

As can be noted from Figure 13c for this category, δu=(1+ε)δl\delta u=(1+\varepsilon)\delta l. With (3.12) we have

δu=(1+ε)(l˙(a;ρ)δa+12l¨(a;ρ)(δa)2+o((δa)2)).\delta{u}=(1+\varepsilon)\left(\dot{l}({a};\rho)\delta{a}+\frac{1}{2}\ddot{l}({a};\rho)(\delta{a})^{2}+o\left((\delta{a})^{2}\right)\right). (3.43)

It follows from (3.43), (3.8), and (3.18a) that

δE1(a;ρ,F)\displaystyle\delta E_{1}(a\,;\rho,F) :=(12F2+2Fw)l˙(a;ρ),\displaystyle:=-\left(\frac{1}{2}{F}^{2}+2F-{w}\right)\dot{l}({a};\rho), (3.44a)
δE2(a;ρ,F)\displaystyle\delta E_{2}(a\,;\rho,F) :=12(F22+2Fw)l¨(a;ρ).\displaystyle:=-\frac{1}{2}\left(\frac{{F}^{2}}{2}+2F-{w}\right)\ddot{l}({a};\rho). (3.44b)
Refer to caption
Figure 8: Forward peeling (a–c) and backward peeling (d–f) on a sinusoidal surface. The function ϱ\varrho of surface profile is the same as the one considered in Figure 4a. (a) and (d) show the numerically calculated true peeling angle ψ\psi, (b) and (e) show the F{F}a{a} plot, and (c) and (f) show representative peeling configurations corresponding to the equilibrium states marked in (b) and (e), respectively. In (b) and (e), the stable, neutral, and unstable equilibrium state are marked with a solid dot, circled dot, and circle, respectively. The yellow regions indicate the occurrence of contact during peeling, while white regions indicate no contact. For the forward (resp. backward) peeling, the peel-off force F+=0.1F^{+}=0.1 (resp. F+=0.03F^{+}=0.03), which is much greater than the peel-off force, 0.033 (resp. 0.003), for peeling on a flat surface with the same nominal peeling angle.
Refer to caption
Figure 9: Forward peeling (a–c) and backward peeling (d–f) on a complicated surface. The function ϱ\varrho of surface profile is the same as the one considered in Figure 4d. (a) and (d) show the numerically calculated true peeling angle ψ\psi, (b) and (e) show the F{F}a{a} plot, and (c) and (f) show representative peeling configurations corresponding to the equilibrium states marked in (b) and (e), respectively. The stable, neutral, and unstable equilibrium states are marked with a solid dot, circled dot, and circle, respectively. The yellow regions indicate the occurrence of contact during peeling, while white regions indicate no contact. For the forward (resp. backward) peeling, the peel-off force F+=0.1F^{+}=0.1 (resp. F+=0.053F^{+}=0.053), which is much greater than the peel-off force, 0.033 (resp. 0.003), for peeling on a flat surface with the same nominal peeling angle.

3.3.4 Remarks on peeling with contact

The results for peeling involving contact shown in Figures 8 and 9 only apply to the sample surfaces shown in Figures 4a and d, respectively. That is, we do not have a general, closed-form, analytical theory for the case in which at least some configurations involve contact. However, we can make the following general, interesting, remarks with regard to the case involving contact.

Theorem 3.1 also holds when the global compatibility condition (3.10) is violated.

Theorem 3.2.

During forward peeling when the global compatibility condition (3.10) is violated, specifically when θ<tan1(ρ˙)\theta<-\tan^{-1}\left(\dot{\rho}^{-}\right), the peel-off force achieves its upper bound, which is (2w)1/2(2w)^{1/2}.

Proof.

Consider the configuration in which the de-adhered length a=aa=a^{-}. The length aa^{-} is defined in (3.10c). Using Algorithm 1 we show that the true peeling angle in this configuration, ψ(a)\psi(a^{-}), is naught.

We start by showing that the aa^{-} configuration involves contact. Let us assume that the aa^{-} configuration does not involve contact. It then follows from Algorithm 1, line 4 that Δρ(x1)>0\Delta\rho(x_{1})>0 for all x<ax<a^{-}. Recalling Δρ\Delta\rho’s definition this last implication can be written more explicitly as

ρ(a)ρ(x1)ax1>tan(θ),\frac{\rho(a)-\rho(x_{1})}{a-x_{1}}>-\tan(\theta), (3.45)

for all x1<ax_{1}<a^{-}. Taking the limit x1ax_{1}\nearrow a^{-} in (3.45) and noting that ϱ˙\dot{\varrho} is a continuous function we get that

ρ˙(a)tan(θ).\dot{\rho}(a^{-})\geq-\tan(\theta). (3.46)

Since ρ˙(a)=:ρ˙\dot{\rho}(a^{-})=:\dot{\rho}^{-}, the left hand side in (3.46) simplifies to ρ˙\dot{\rho}^{-}. It follows from our hypothesis that the global compatibility condition is violated that the right hand side of (3.46) is greater than ρ˙\dot{\rho}^{-}. Thus, we get a contradiction. Hence, our assumption that the aa^{-} configuration does not involve contact is false.

Since the aa^{-} configuration involves contact we need to move to line 7 of Algorithm 1 for determining the configuration’s true peeling angle. Noting that we have assumed ρ¨\ddot{\rho} to be a continuous function and, from (3.10b), that ρ˙\dot{\rho}^{-} is ρ˙\dot{\rho}’s minimum value we get that ρ¨(a)=0\ddot{\rho}(a^{-})=0. This last result together with (3.13a) and (3.35) implies that k(a)=0k(a^{-})=0. The function ϱ\varrho’s property that it is a surjective function with the range [1,1][-1,1] implies that ρ˙\dot{\rho}^{-} is negative. As a consequence of these last two implications we need to move to line 8 from line 7 in the algorithm. Since in forward peeling θ<π/2\theta<\pi/2 we get from line 8 that the true peeling angle in the aa^{-} configuration is naught, i.e., ψ(a)=0\psi(a^{-})=0.

In §3.3 we discussed that the equilibrium force corresponding to a configuration with de-adhered length aa is (ψ(𝒶))\mathpzc{F}(\psi(a)). Thus, we the aa^{-} configuration’s equilibrium force is (0)\mathpzc{F}(0), which simplifies to (2w)1/2(2w)^{1/2}.

Recall that (2w)1/2(2w)^{1/2} is the function \mathpzc{F}’s maximum value. This fact in conjunction with F+F^{+}’s definition (3.22) and the final result from the previous paragraph imply that F+=(2w)1/2F^{+}=(2w)^{1/2}. ∎

Theorem 3.3.

During backward peeling when the global compatibility condition (3.10) is violated, specifically when θ>πtan1(ρ˙+)\theta>\pi-\tan^{-1}\left(\dot{\rho}^{+}\right), the peel-initiation force achieves its lower bound, which is (4+2w)1/22(4+2w)^{1/2}-2.

We omit our proof for Theorem 3.3. Since it is quite similar to the one we provided for Theorem 3.2, except that in it we focus on the configuration with de-adhered length a+a^{+} instead of the configuration with de-adhered length aa^{-}.

Theorem 3.4.

During backward peeling when the global compatibility condition is violated, specifically when

θ\displaystyle\theta >πtan1(ρ˙+)\displaystyle>\pi-\tan^{-1}\left(\dot{\rho}^{+}\right) (3.47a)
the equilibrium force is always less than or equal to (ψlb)\mathcal{F}\left(\psi_{\rm lb}^{-}\right), where
ψlb\displaystyle\psi_{\rm lb}^{-} :=π+tan1(ρ˙)tan1(ρ˙+).\displaystyle:=\pi+\tan^{-1}\left(\dot{\rho}^{-}\right)-\tan^{-1}\left(\dot{\rho}^{+}\right). (3.47b)
Proof.

In §3.3 we discovered that during backward peeling when the global compatibility condition is violated there will exist some configurations that involve contact and others that do not (see Figures 8e and 9e.)

When a configuration involves contact the true peeling angle is given by (3.15). Since, by hypothesis, θ>πtan1(ρ˙+)\theta>\pi-\tan^{-1}\left(\dot{\rho}^{+}\right) and, since ρ˙(a)ρ˙\dot{\rho}(a)\geq\dot{\rho}^{-} by definition, tan1(ρ˙(a))tan1(ρ˙)\tan^{-1}\left(\dot{\rho}(a)\right)\geq\tan^{-1}\left(\dot{\rho}^{-}\right) it follows from (3.15) that during backward peeling when the configuration does not involve contact the true peeling angle is greater than π+tan1(ρ˙)tan1(ρ˙+)\pi+\tan^{-1}\left(\dot{\rho}^{-}\right)-\tan^{-1}\left(\dot{\rho}^{+}\right), which is nothing but ψlb\psi_{\rm lb}^{-}.

When a configuration involves contact then it follows from Algorithm 1 that the true peeling angle is either equal to π\pi (local contact) or to π+tan1(ρ˙(a))tan1(ρ˙(c))\pi+\tan^{-1}\left(\dot{\rho}(a)\right)-\tan^{-1}\left(\dot{\rho}(c)\right) (non-local contact), where recall that aa and cc are the abscissae of the peeling and contact fronts, respectively. Since we have assumed ρ\rho to be a non-constant function it follows that ρ˙+>ρ˙\dot{\rho}^{+}>\dot{\rho}^{-} and hence that π>ψlb\pi>\psi_{\rm lb}^{-}. It follows from the definitions of ρ˙+\dot{\rho}^{+}, ρ˙\dot{\rho}^{-}, and the monotonicity of tan1\tan^{-1} that π+tan1(ρ˙(a))tan1(ρ˙(c))ψlb\pi+\tan^{-1}\left(\dot{\rho}(a)\right)-\tan^{-1}\left(\dot{\rho}(c)\right)\geq\psi_{\rm lb}^{-}. Thus, in the case of contact the true peeling angle is greater than or equal to ψlb\psi_{\rm lb}^{-}.

The deductions in the last two paragraphs can be summarized by saying that when (3.47a) holds the true peeling angle is always greater than or equal to ψlb\psi_{\rm lb}^{-}. In §3.3 we discussed that the equilibrium force corresponding to the true peeling angle ψ(a)\psi(a) is always (ψ(𝒶))\mathpzc{F}(\psi(a)), irrespective of whether or not the configuration involves contact. Since \mathpzc{F} is a monotonically decreasing function the last two statements imply that under (3.47a) the equilibrium peeling force is always less than or equal to (ψlb)=:Fub+\mathcal{F}\left(\psi_{\rm lb}^{-}\right)=:F_{\rm ub}^{+}. ∎

For illustrating Theorem 3.3, we mark ψlb\psi_{\rm lb}^{-} and Fub+{F}_{\rm ub}^{+} for the case of backward peeling under (3.47a) on a simple wavy surface in Figures 8d and e, respectively, and on a complicated wavy surface in Figures 9d and e, respectively.

4 Angle-independent optimal peel-off force

In this section, we analyze the asymptotic value of the peel-off force F+F^{+} when the substrate’s aspect ratio α\alpha, or its root-mean-square (RMS)999The RMS roughness of the substrate’s surface is equal to α(01ϱ(x1)2𝑑x1)1/2\alpha\left(\int_{0}^{1}\varrho(x_{1})^{2}\,dx_{1}\right)^{1/2}. roughness, becomes large. This is equivalent to, e.g., the case where the substrate’s surface’s periodicity λ\lambda becomes vanishingly small in comparison to its amplitude AA.

It follows from Algorithm 1 that as α\alpha becomes large all configurations become contact configurations, irrespective of aa or ϱ\varrho, if θ\theta is different from π/2\pi/2; and if θ=π/2\theta=\pi/2 none of the configurations involve contact.

When θ=π/2\theta=\pi/2, since none of the configurations involve contact, we can use the results given in §3.2. Specifically, using (3.23b) we get that as α\alpha becomes large ψ\psi^{-} becomes vanishingly small, since in that limit ρ˙\dot{\rho}^{-} tends to negative infinity. Taking the limit ψ0\psi^{-}\to 0 in (3.23a) we get the result that F+F^{+} approaches its upper bound (2w)1/2(2w)^{1/2} as α\alpha becomes large. This result is shown illustrated in Figure 10.

When θ<π/2\theta<\pi/2, the condition θ<tan1(ρ˙)\theta<-\tan^{-1}(\dot{\rho}^{-}) will inevitably get violated for a large enough α\alpha. Thus, from Theorem 3.2 we get that F+F^{+} will eventually equal its upper bound of (2w)1/2(2w)^{1/2} as α\alpha becomes large.

When θ>π/2\theta>\pi/2, let us choose an aa for which ϱ˙(a)\dot{\varrho}(a) is negative. Now consider a sequence of configurations that all have that same aa, ϱ\varrho, and θ>π/2\theta>\pi/2 but increasingly larger values of α\alpha. It follows from Algorithm 1 that when α\alpha becomes large enough all subsequent configurations will be of the non-local-contact type, and that the true peeling angle in all of them can be computed as π+tan1(ρ˙(a))tan1(ρ˙(c))\pi+\tan^{-1}\left(\dot{\rho}(a)\right)-\tan^{-1}\left(\dot{\rho}(c)\right). Recall that cc is the contact front’s abscissa. We would expect cc to vary between the configurations. However, it can be shown, again using Algorithm 1, that once the configurations become of the non-local-contact type the cc in them also remains fixed, and furthermore that the value of ϱ˙\dot{\varrho} at that cc is positive. Since ρ˙=αϱ˙\dot{\rho}=\alpha\dot{\varrho} and ϱ˙\dot{\varrho} is negative at aa and positive at cc as α\alpha becomes large tan1(ρ˙(a))\tan^{-1}\left(\dot{\rho}\left(a\right)\right) and tan1(ρ˙(c))\tan^{-1}\left(\dot{\rho}\left(c\right)\right) tend to π/2\mp\pi/2, respectively, and, consequently, the true peeling angles tend to naught. Recall from the discussion in §3.3 that irrespective of whether or not a configuration involves contact, the equilibrium force in that configuration can always be computed as (ψ(a))\mathcal{F}(\psi(a)), where ψ(a)\psi(a) is the configuration’s true peeling angle. Thus, as α\alpha becomes large the equilibrium force in the sequence converges to (0)=(2w)1/2\mathcal{F}(0)=(2w)^{1/2}. From (3.22) we know that the peel-off force for each geometry corresponding to a configuration in the sequence, namely that defined by the profile αϱ\alpha\varrho and the peeling angle θ\theta, is greater than or equal to that configuration’s equilibrium force. As we noted in the discussion immediately ensuing (3.22) the peel-off force, irrespective of profile or peeling-angle, is bounded above by (2w)1/2(2w)^{1/2}. It follows from the last three statements that for any fixed ϱ\varrho and θ>π/2\theta>\pi/2 as α\alpha becomes large the peel-off force tends to its upper bound (2w)1/2(2w)^{1/2}.

Refer to caption
Figure 10: The plot of peel-off force F+F^{+} as a function of nominal peeling angle θ\theta for the peeling on (a) simple and (b) complicated surfaces for a series of increasing surface’s aspect ratio α\alpha. The surface profiles, ϱ\varrho, are the same as those considered in Figures 4c and f, respectively.

In summary, from the previous three paragraphs we have the important conclusion that as the surface roughness, α\alpha, is increased the peel-off force F+F^{+} tends to its upper bound (2w)1/2(2w)^{1/2}. This happens independent of the surface’s shape, ϱ\varrho, and the peeling angle, θ\theta. We call this phenomenon angle-independent optimal adhesion. This phenomenon is quite interesting considering that adhesion on a flat surface is highly dependent on the peeling angle, and the optimal adhesion is only attained at a single peeling angle, namely for θ=0\theta=0.

We numerically computed the peel-off force F+F^{+} for the simple and complicated wavy surface shapes, which we first considered in §3.2, for various peeling angles. For each surface shape and peeling angle we calculated F+F^{+} for a sequence of geometries of increasing α\alpha values. The results of our calculations are shown in Figure 10. As can be seen, at small α\alpha values, for instance α=1.0\alpha=1.0 and 5.05.0, the peel-off force F+{F}^{+} depends strongly on θ\theta. However, as α\alpha increases the dependence of F+{F}^{+} on θ\theta becomes weak, such that at large α\alpha values, e.g. α=10.0\alpha=10.0, the calculated peel-off forces appear to be essentially independent of the peeling angle. Finally, irrespective of the peeling angle or the surface shape, the calculated peel-off force values appears to approach (2w)1/2(2w)^{1/2} from below as α\alpha increases.

5 Concluding remarks

We conclude this paper by briefly commenting on the effect of bending strain energy on the peel-off force.

Peng and Chen [42] investigated the peeling of an elastic film on a sinusoidal surface. Their sinusoidal surface is the same our simple wavy surface that we first considered in §3.2, and is shown illustrated in Figure 4a. In their analysis they took into account the film’s bending energy, where as in our model we only consider the strain energy due to tension and ignore the strain energy due to bending. They do not consider contact between the thin-film and the substrate in their analysis. Therefore, some insight into the effect of the bending strain energy can be garnered by comparing their analysis to the results that we present in §3.2 (peeling with no contact) when they are particularized to the case of simple wavy surface.

In Peng and Chen’s model the equilibrium peeling force, FF, is related to the de-adhered length, aa, as

F(a)=cos(ψ(a))1+((cos(ψ(a))1)2+2w2π4α2h¯2(1+cos(4πa))3(1+4π2α2sin2(2πa))1/2)1/2,F(a)=\cos(\psi({a}))-1+\left(\left(\cos(\psi({a}))-1\right)^{2}+2{w}-\frac{2\pi^{4}\alpha^{2}\bar{h}^{2}(1+\cos(4\pi{a}))}{3(1+4\pi^{2}\alpha^{2}\sin^{2}(2\pi{a}))^{1/2}}\right)^{1/2}, (5.1)

where h¯=h/λ\bar{h}=h/\lambda. On particularizing the results in §3.2 to the case of simple wavy surface our model predicts the same relation between the equilibrium force and de-adhered length as (5.1) except that in it there are no terms containing h¯\bar{h}. That is, in our model’s prediction the third terms from the left within the parenthesis of (5.1), containing h¯\bar{h}, is absent. Since the third term scales with h¯\bar{h} as h¯2\bar{h}^{2}, Peng and Chen’s results converge to our results as h¯0\bar{h}\to 0. This fact can also be noted from Figure 11, in which we compare the predictions from Peng and Chen’s model with those from our model for the case of no-contact and peeling on a simple wavy surface for various h¯\bar{h} (in subfigure a) and θ\theta (in subfigure (b)) values. Even though the above comparison is only for a particular substrate profile, namely the simple wavy surface, we believe that it is reasonable to expect that the effect of the bending strain energy on the equilibrium forces will decrease with decreasing film thicknesses.

Another interesting feature of the above above comaparision that is revealed by the Figure 11 and merits further investigation is that ignoring the bending strain energy seems to have no affect on the peel-off force. The peel-off force F+F^{+} which we might recall is the supremum of all equilibrium forces on a given substrate profile and peeling angle. Specifically, as per Figure 11 the surpremum of the equilibrium forces predicted by our’s as well as Peng and Chen’s model appear to be the same. This numerical evidence prompts us conjecture that the bending strain energy, or the film thickness, will have no effect on a film’s peel-off force irrespective of the substrate profile or peeling angle.

Refer to caption
Figure 11: (a) The F{F}a{a} plot for different values of thin film thicknesses h¯\bar{h} that considers the thin elastic film’s bending energy for peeling on a sinusoidal surface at θ=π/3\theta=\pi/3. (b) The F{F}a{a} plot for different values of nominal peeling angle θ\theta for peeling a thin film with thickness h¯=0\bar{h}=0 (red curves) and h¯=0.06\bar{h}=0.06 (black curves) on a sinusoidal surface. For the sinusoidal surface in (a) and (b), we take A=0.25A=0.25, λ=1.0\lambda=1.0, ϱ(x1)=cos(2πx1)\varrho(x_{1})=-\cos(2\pi x_{1}), and w=0.005{w}=0.005. For h¯=0\bar{h}=0, the stable, neutral, and unstable equilibrium state are marked with a solid dot, circled dot, and circle, respectively, from the stability analysis of our model.

Acknowledgment

The authors gratefully acknowledge support from the Office of Naval Research (Dr. Timothy Bentley) [Panther Program, grant number N000141812494] and the National Science Foundation [Mechanics of Materials and Structures Program, grant number 1562656]. Weilin Deng is partially supported by a graduate fellowship from the China Scholarship Council.

Appendix A Derivation of δu\delta u for peeling involving contact

Refer to caption
Figure 12: (a) The schematic of forward peeling involving contact that considers the peeling front PP moves to P~\tilde{P} after an infinitesimal perturbation. The contact front accordingly changes from CC to C~\tilde{C}. (b) and (c) show non-local and local type contact, respectively.
Refer to caption
Figure 13: (a) The schematic of backward peeling involving contact that considers the peeling front PP moves to P~\tilde{P} after an infinitesimal perturbation. The contact front accordingly changes from CC to C~\tilde{C}. (b) and (c) show non-local and local type contact, respectively.

Configurations that involve contact during a generic forward peeling experiment are shown in Figure 12, and similar configurations from a backward peeling experiment are shown in Figure 13. For concreteness we focus on the forward peeling experiment, however, our remarks apply to the backward peeling experiment as well. The peeled part and the remainder of the adhered part of the thin film in the configuration 𝜿\bm{\kappa} are shown, respectively, in dark orange and blue in Figure 12a. We perturb 𝜿\bm{\kappa} slightly to obtain the nearby configuration 𝜿~\tilde{\bm{\kappa}}. The peeled part of the thin film in 𝜿~\tilde{\bm{\kappa}} is shown in light orange in Figure 12a. Recall that PP and CC denote the peeling and contact fronts in 𝜿\bm{\kappa}. We denote the peeling and contact fronts in 𝜿~\tilde{\bm{\kappa}} as P~\tilde{P} and C~\tilde{C}.

The quantity that we aim to compute, namely, δu\delta u, is related to the thin film’s kinematics. The thin film’s kinematics take place in \mathcal{E}. However, since the thin film’s kinematics do not vary in the \mathbscre3\mathbscr{e}_{3} direction, δu\delta u can be computed by only analyzing the kinematics that take place in, say, \mathcal{E}’s x1x_{1}-x2x_{2} plane. Therefore, we will be focusing all our analysis related to computing δu\delta u only to \mathcal{E}’s x1x_{1}-x2x_{2} plane. In order to avoid the introduction of more symbols, we will use the same symbols that we introduced to refer to subsets of \mathcal{E} to refer to the quantities that result from the projection of those subsets into the x1x_{1}-x2x_{2} plane. For instance. The peeling and contact fronts, PP and CC, are line segments in \mathcal{E}. However, we will be denoting their projections into the x1x_{1}x2x_{2} plane, which are points in \mathcal{E}, also as PP and CC, and refer to them as the peeling and contact point, respectively. Similarly, we denote the peeling and contact points in 𝜿~\tilde{\bm{\kappa}} as P~\tilde{P} and C~\tilde{C}.

We consider a general contact scenario in which the contact region Γc𝒟\Gamma_{c}\in\mathcal{D} consists of several disjoint contact patches, Γc1\Gamma_{c_{1}},Γc2\Gamma_{c_{2}}, etc. In Figure 12a we mark the regions on the substrate which mate with these contact patches in 𝜿\bm{\kappa} as 𝒙(Γc1)\bm{x}\left(\Gamma_{\rm c_{1}}\right), 𝒙(Γc2)\bm{x}\left(\Gamma_{\rm c_{2}}\right),…, 𝒙(Γcn)\bm{x}\left(\Gamma_{\rm c_{n}}\right). We call 𝒙(Γc1)\bm{x}\left(\Gamma_{\rm c_{1}}\right) the first contact region, and 𝒙(Γcn)\bm{x}\left(\Gamma_{\rm c_{n}}\right) the last contact region. By definition, the point on the substrate where the first contact region begins is the contact point CC. We mark the point where the first contact region ends as DD. Similarly, we mark the points where the ithi^{\rm th}, i=2,,ni=2,\ldots,n, contact region begins and ends as CiC_{i} and DiD_{i}, respectively. We denote the location of the thin film’s terminal end in 𝜿\bm{\kappa} as XX.

As we perturb the thin film’s configuration from 𝜿\bm{\kappa} to 𝜿~\tilde{\bm{\kappa}}, the peeling front moves from PP to P~\tilde{P}, the contact front moves from CC to C~\tilde{C}, and the terminal end from XX to X~\tilde{X}. Interestingly, however, none of the other features that define the thin film’s geometry move during this perturbation. Specifically, in 𝜿~\tilde{\bm{\kappa}} the first contact region still ends at DD, and all the other contact regions still begin and end at the same points that they did in 𝜿\bm{\kappa}.

The length of the peeled part of the film in 𝜿\bm{\kappa} can be computed as the sum of the lengths of the line segment \macc@depthΔ\frozen@everymath\macc@group\macc@set@skewchar\macc@nested@a111C\macc@depth\char 1\relax\frozen@everymath{\macc@group}\macc@set@skewchar\macc@nested@a 111{C}, the arc ¿ CD\textstyle CD , the line segment \macc@depthΔ\frozen@everymath\macc@group\macc@set@skewchar\macc@nested@a111D\macc@depth\char 1\relax\frozen@everymath{\macc@group}\macc@set@skewchar\macc@nested@a 111{D}, the arcs ¿ CiDi\textstyle C_{i}D_{i} , i=2,,ni=2,\ldots,n, and the line segments \macc@depthΔ\frozen@everymath\macc@group\macc@set@skewchar\macc@nested@a111D\macc@depth\char 1\relax\frozen@everymath{\macc@group}\macc@set@skewchar\macc@nested@a 111{D}, i=2,,n1i=2,\ldots,n-1, and, finally, the line segment \macc@depthΔ\frozen@everymath\macc@group\macc@set@skewchar\macc@nested@a111D\macc@depth\char 1\relax\frozen@everymath{\macc@group}\macc@set@skewchar\macc@nested@a 111{D}. Based on the discussion in the previous paragraph, the length of the peeled part in 𝜿~\tilde{\bm{\kappa}} is equal to the sum of the lengths of the line segment \macc@depthΔ\frozen@everymath\macc@group\macc@set@skewchar\macc@nested@a111\macc@depth\char 1\relax\frozen@everymath{\macc@group}\macc@set@skewchar\macc@nested@a 111{}, the arc ¿ C~D\textstyle\tilde{C}D , and, as before, the line segment \macc@depthΔ\frozen@everymath\macc@group\macc@set@skewchar\macc@nested@a111D\macc@depth\char 1\relax\frozen@everymath{\macc@group}\macc@set@skewchar\macc@nested@a 111{D}, the arcs ¿ CiDi\textstyle C_{i}D_{i} , and the line segments \macc@depthΔ\frozen@everymath\macc@group\macc@set@skewchar\macc@nested@a111D\macc@depth\char 1\relax\frozen@everymath{\macc@group}\macc@set@skewchar\macc@nested@a 111{D}, and, finally, the line segment DnX~¯\overline{D_{n}\tilde{X}}. The difference in the length of the peeled part between 𝜿\bm{\kappa} and 𝜿~\tilde{\bm{\kappa}}, therefore, is

DnX~¯DnX¯+ ¿ ~CD ¿ CD +P~C~¯PC¯.\overline{D_{n}\tilde{X}}-\overline{D_{n}X}+\mathchoice{\vbox{ \hbox{\leavevmode\resizebox{14.11252pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\displaystyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{14.11252pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\textstyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{11.54543pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptstyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{9.83403pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptscriptstyle\tilde{C}D$}}}-\mathchoice{\vbox{ \hbox{\leavevmode\resizebox{16.41943pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\displaystyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{16.41943pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\textstyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{11.49359pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptstyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{8.2097pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptscriptstyle CD$}}}+\overline{\tilde{P}\tilde{C}}-\overline{PC}.

The length difference can, alternately, also be computed as δl(1+ε)\delta l(1+\varepsilon). Equating these two expressions for the length difference and noting that DnX~¯DnX¯\overline{D_{n}\tilde{X}}-\overline{D_{n}X} is in fact δu\delta u we get that

δu=δl(1+ε)( ¿ ~CD ¿ CD )P~C~¯+PC¯.\delta u=\delta l(1+\varepsilon)-\left(\mathchoice{\vbox{ \hbox{\leavevmode\resizebox{14.11252pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\displaystyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{14.11252pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\textstyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{11.54543pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptstyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{9.83403pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptscriptstyle\tilde{C}D$}}}-\mathchoice{\vbox{ \hbox{\leavevmode\resizebox{16.41943pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\displaystyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{16.41943pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\textstyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{11.49359pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptstyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{8.2097pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptscriptstyle CD$}}}\right)-\overline{\tilde{P}\tilde{C}}+\overline{PC}. (A.1)

The term ¿ ~CD ¿ CD \mathchoice{\vbox{ \hbox{\leavevmode\resizebox{14.11252pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\displaystyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{14.11252pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\textstyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{11.54543pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptstyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{9.83403pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptscriptstyle\tilde{C}D$}}}-\mathchoice{\vbox{ \hbox{\leavevmode\resizebox{16.41943pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\displaystyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{16.41943pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\textstyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{11.49359pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptstyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{8.2097pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptscriptstyle CD$}}} in (A.1) can be computed as

¿ ~CD ¿ CD =±cc~(1+ρ˙(x1)2)1/2𝑑x1,\mathchoice{\vbox{ \hbox{\leavevmode\resizebox{14.11252pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\displaystyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{14.11252pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\textstyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{11.54543pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptstyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{9.83403pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptscriptstyle\tilde{C}D$}}}-\mathchoice{\vbox{ \hbox{\leavevmode\resizebox{16.41943pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\displaystyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{16.41943pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\textstyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{11.49359pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptstyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{8.2097pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptscriptstyle CD$}}}=\pm\int^{\tilde{c}}_{c}\left(1+\dot{\rho}(x_{1})^{2}\right)^{1/2}\,dx_{1}, (A.2)

where the plus sign is for the case of forward peeling, while the minus sign is for the case of backward peeling.

The points PP, CC, P~\tilde{P}, and C~\tilde{C} are shown marked in, e.g., Figure 12a. From the definitions of peeling and contact fronts it follows that the coordinates of the points PP and CC are, respectively, (a,ρ(a))(a,\rho(a)) and (c,ρ(c))(c,\rho(c)). By denoting the abscissa of the points P~\tilde{P} and C~\tilde{C} as a~\tilde{a} and c~\tilde{c} it follows for similar reasons that these points’ coordinates are (a~,ρ(a~))(\tilde{a},\rho(\tilde{a})) and (c~,ρ(c~)(\tilde{c},\rho(\tilde{c}), respectively. The line segments \macc@depthΔ\frozen@everymath\macc@group\macc@set@skewchar\macc@nested@a111P\macc@depth\char 1\relax\frozen@everymath{\macc@group}\macc@set@skewchar\macc@nested@a 111{P} and \macc@depthΔ\frozen@everymath\macc@group\macc@set@skewchar\macc@nested@a111\macc@depth\char 1\relax\frozen@everymath{\macc@group}\macc@set@skewchar\macc@nested@a 111{} are tangent to the graph of ρ\rho at CC and C~\tilde{C}, respectively. (These aspects of the film’s geometry are especially clear in Figure 12b.) Using this information it can be shown that

ρ(a)ρ(c)\displaystyle\rho\left(a\right)-\rho\left(c\right) =ρ˙(c)(ac),\displaystyle=\dot{\rho}\left(c\right)(a-c), (A.3a)
ρ(a~)ρ(c~)\displaystyle\rho\left(\tilde{a}\right)-\rho\left(\tilde{c}\right) =ρ˙(c~)(a~c~).\displaystyle=\dot{\rho}\left(\tilde{c}\right)(\tilde{a}-\tilde{c}). (A.3b)

Using the knowledge of PP, CC, P~\tilde{P}, and C~\tilde{C}’s coordinates, and (A.3) it can be shown that

PC¯\displaystyle\overline{PC} =(1+ρ˙(c)2)1/2|ca|,\displaystyle=\left(1+\dot{\rho}(c)^{2}\right)^{1/2}\lvert c-a\rvert, (A.4)
P~C~¯\displaystyle\overline{\tilde{P}\tilde{C}} =(1+ρ˙(c~)2)1/2|c~a~|.\displaystyle=\left(1+\dot{\rho}(\tilde{c})^{2}\right)^{1/2}\lvert\tilde{c}-\tilde{a}\rvert. (A.5)

Substituting ¿ ~CD ¿ CD \mathchoice{\vbox{ \hbox{\leavevmode\resizebox{14.11252pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\displaystyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{14.11252pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\textstyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{11.54543pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptstyle\tilde{C}D$}}}{\vbox{ \hbox{\leavevmode\resizebox{9.83403pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptscriptstyle\tilde{C}D$}}}-\mathchoice{\vbox{ \hbox{\leavevmode\resizebox{16.41943pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\displaystyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{16.41943pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\textstyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{11.49359pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptstyle CD$}}}{\vbox{ \hbox{\leavevmode\resizebox{8.2097pt}{0.0pt}{{\char 62\relax}}} \nointerlineskip\hbox{\set@color$\scriptscriptstyle CD$}}}, PC¯\overline{PC}, and P~C~¯\overline{\tilde{P}\tilde{C}} in (A.1) with the right hand sides of (A.2), (A.4), and (A.5), respectively; then writing a~\tilde{a} as a+δaa+\delta a, and cc as the right hand side of (A.9) in the resulting equation; and then, finally, expanding the resulting equation in series of δa\delta a, we get that as δa0\delta a\to 0

δu=\displaystyle\delta{u}= (1+εcos(ψ(a)))l˙(a;ρ)δa\displaystyle\left(1+\varepsilon-\cos(\psi(a))\right)\dot{l}({a};\rho)\delta{a} (A.6)
+12((1+ε)l¨(a;ρ)l˙(a;ρ)2sin2(ψ(a))+(sin(ψ(a))cos(ψ(a))ρ˙(a))ρ¨(a)l˙(a;ρ))(δa)2+o((δa)2),\displaystyle+\frac{1}{2}\left((1+\varepsilon)\ddot{l}({a};\rho)-\frac{\dot{l}({a};\rho)^{2}}{\ell}\sin^{2}(\psi(a))+\frac{\left(\sin(\psi(a))-\cos(\psi(a))\dot{\rho}(a)\right)\ddot{\rho}(a)}{\dot{l}(a;\rho)}\right)(\delta{a})^{2}+o\left((\delta a)^{2}\right),

where \ell is an alias for PC¯\overline{PC}. We introduce this new symbol so as to make some of the results that derive from (A.6) appear more compact. The result (A.6) applies to both forward as well as backward peeling. In arriving at (A.6) we used (3.13), and the result that

cos(ψ(a))=±1+ρ˙(a)ρ˙(c)((1+ρ˙(a)2)(1+ρ˙(c)2))12,\cos\left(\psi(a)\right)=\pm\frac{1+\dot{\rho}(a)\dot{\rho}(c)}{\left(\left(1+\dot{\rho}(a)^{2}\right)\left(1+\dot{\rho}(c)^{2}\right)\right)^{\frac{1}{2}}}, (A.7)

where the plus sign is for the case of forward peeling, while the minus sign is for the case of backward peeling. The result (A.7) follows from Algorithm 1.

A.1 The asymptotic behavior of δc\delta c as δa0\delta a\to 0

Expressing a~\tilde{a} as a+δaa+\delta a and c~\tilde{c} as c+f(δa)c+f\left(\delta a\right), where ff is some real valued analytic function over \mathbb{R}, in (A.3b), and then expanding both sides of the resulting equation in series of δa\delta a, we get that as δa0\delta a\to 0

((ca)ρ˙(c)+ρ(a)ρ(c))+((ca)f˙(0)ρ¨(c)+ρ˙(a)ρ˙(c))δa+o(δa)=0.\left((c-a)\dot{\rho}(c)+\rho(a)-\rho(c)\right)+\left((c-a)\dot{f}(0)\ddot{\rho}(c)+\dot{\rho}(a)-\dot{\rho}(c)\right)\delta a+o\left(\delta a\right)=0. (A.8)

In arriving at (A.8) we made use of the identity that f(0)=0f(0)=0, which is a consequence of the requirement that c~c\tilde{c}\to c as δa0\delta a\to 0. It follows from (A.3a) and (A.8) that

δc=ρ˙(a)ρ˙(c)ρ¨(c)(ac)δa+o(δa).\delta c=\frac{\dot{\rho}(a)-\dot{\rho}(c)}{\ddot{\rho}(c)(a-c)}\delta a+o\left(\delta a\right). (A.9)

References

  • [1] Wilhelm Barthlott and Christoph Neinhuis. Purity of the sacred lotus, or escape from contamination in biological surfaces. Planta, 202(1):1–8, 1997.
  • [2] Li Wen, James C Weaver, and George V Lauder. Biomimetic shark skin: design, fabrication and hydrodynamic function. Journal of Experimental Biology, 217(10):1656–1666, 2014.
  • [3] KNG Fuller and David Tabor. The effect of surface roughness on the adhesion of elastic solids. Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences, 345(1642):327–342, 1975.
  • [4] John M Levins and T Kyle Vanderlick. Impact of roughness on the deformation and adhesion of a rough metal and smooth mica in contact. The Journal of Physical Chemistry, 99(14):5067–5076, 1995.
  • [5] RA Quon, RF Knarr, and TK Vanderlick. Measurement of the deformation and adhesion of rough solids in contact. The Journal of Physical Chemistry B, 103(25):5320–5327, 1999.
  • [6] Yakov I Rabinovich, Joshua J Adler, Ali Ata, Rajiv K Singh, and Brij M Moudgil. Adhesion between nanoscale rough surfaces: Ii. measurement and comparison with theory. Journal of Colloid and Interface Science, 232(1):17–24, 2000.
  • [7] GAD Briggs and BJ Briscoe. The effect of surface topography on the adhesion of elastic solids. Journal of Physics D: Applied Physics, 10(18):2453, 1977.
  • [8] Haneesh Kesari, Joseph C Doll, Beth L Pruitt, Wei Cai, and Adrian J Lew. Role of surface roughness in hysteresis during adhesive elastic contact. Philosophical Magazine & Philosophical Magazine Letters, 90(12):891–902, 2010.
  • [9] Haneesh Kesari and Adrian J Lew. Effective macroscopic adhesive contact behavior induced by small surface roughness. Journal of the Mechanics and Physics of Solids, 59(12):2488–2510, 2011.
  • [10] Weilin Deng and Haneesh Kesari. Depth-dependent hysteresis in adhesive elastic contacts at large surface roughness. Scientific Reports, 9(1):1639, 2019.
  • [11] Weilin Deng and Haneesh Kesari. Effect of machine stiffness on interpreting contact force–indentation depth curves in adhesive elastic contact experiments. Journal of the Mechanics and Physics of Solids, 131:404–423, 2019.
  • [12] Weilin Deng and Haneesh Kesari. Molecular statics study of depth-dependent hysteresis in nano-scale adhesive elastic contacts. Modelling and Simulation in Materials Science and Engineering, 25(5):055002, 2017.
  • [13] Jean-Claude Michel, Hervé Moulinec, and P Suquet. Effective properties of composite materials with periodic microstructure: a computational approach. Computer methods in applied mechanics and engineering, 172(1-4):109–143, 1999.
  • [14] Jerry L Ericksen, David Kinderlehrer, Robert Kohn, and J-L Lions. Homogenization and effective moduli of materials and media, volume 1. Springer Science & Business Media, 2012.
  • [15] Pedro Ponte Castaneda. Second-order homogenization estimates for nonlinear composites incorporating field fluctuations: I–theory. Journal of the Mechanics and Physics of Solids, 50(4):737–757, 2002.
  • [16] Valentin L Popov, Roman Pohrt, and Qiang Li. Strength of adhesive contacts: Influence of contact geometry and material gradients. Friction, 5(3):308–325, 2017.
  • [17] Michele Ciavarella, J Joe, Antonio Papangelo, and JR Barber. The role of adhesion in contact mechanics. Journal of the Royal Society Interface, 16(151):20180738, 2019.
  • [18] R.S. Rivlin. The effective work of adhesion. Paint Technol., 9:215–218, 1944.
  • [19] K. Kendall. Thin-film peeling-the elastic term. Journal of Physics D: Applied Physics, 8(13):1449, 1975.
  • [20] KL Johnson, K Kendall, and AD Roberts. Surface energy and the contact of elastic solids. Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, 324(1558):301–313, 1971.
  • [21] 3d printed helmet.
  • [22] Xue Feng, Matthew A Meitl, Audrey M Bowen, Yonggang Huang, Ralph G Nuzzo, and John A Rogers. Competing fracture in kinetically controlled transfer printing. Langmuir, 23(25):12555–12560, 2007.
  • [23] Noshir S Pesika, Yu Tian, Boxin Zhao, Kenny Rosenberg, Hongbo Zeng, Patricia McGuiggan, Kellar Autumn, and Jacob N Israelachvili. Peel-zone model of tape peeling based on the gecko adhesive system. The Journal of Adhesion, 83(4):383–401, 2007.
  • [24] Bin Chen, Peidong Wu, and Huajian Gao. Pre-tension generates strongly reversible adhesion of a spatula pad on substrate. Journal of The Royal Society Interface, 6(35):529–537, 2008.
  • [25] Alain Molinari and Guruswami Ravichandran. Peeling of elastic tapes: effects of large deformations, pre-straining, and of a peel-zone model. The Journal of Adhesion, 84(12):961–995, 2008.
  • [26] Matthew R Begley, Rachel R Collino, Jacob N Israelachvili, and Robert M McMeeking. Peeling of a tape with large deformations and frictional sliding. Journal of the Mechanics and Physics of Solids, 61(5):1265–1279, 2013.
  • [27] Panayiotis Gialamas, Benjamin Völker, Rachel R Collino, Matthew R Begley, and Robert M McMeeking. Peeling of an elastic membrane tape adhered to a substrate by a uniform cohesive traction. International Journal of Solids and Structures, 51(18):3003–3011, 2014.
  • [28] N. Menga, L. Afferrante, NM Pugno, and G. Carbone. The multiple v-shaped double peeling of elastic thin films from elastic soft substrates. Journal of the Mechanics and Physics of Solids, 113:56–64, 2018.
  • [29] Kyung-Suk Kim and Junglhl Kim. Elasto-plastic analysis of the peel test for thin film adhesion. Journal of Engineering Materials and Technology, 110(3):266–273, 1988.
  • [30] A.J. Kinloch, C.C. Lau, and J.G. Williams. The peeling of flexible laminates. International Journal of Fracture, 66(1):45–70, 1994.
  • [31] Yueguang Wei and John W. Hutchinson. Interface strength, work of adhesion and plasticity in the peel test. International Journal of Fracture, 93(1):315–333, 1998.
  • [32] M.J. Loukis and N. Aravas. The effects of viscoelasticity in the peeling of polymeric films. The Journal of Adhesion, 35(1):7–22, 1991.
  • [33] L. Afferrante and G. Carbone. The ultratough peeling of elastic tapes from viscoelastic substrates. Journal of the Mechanics and Physics of Solids, 96:223–234, 2016.
  • [34] K. Kendall. Control of cracks by interfaces in composites. Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, 341(1627):409–428, 1975.
  • [35] Animangsu Ghatak, L Mahadevan, Jun Young Chung, Manoj K Chaudhury, and Vijay Shenoy. Peeling from a biomimetically patterned thin elastic film. Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, 460(2049):2725–2735, 2004.
  • [36] Jun Young Chung and Manoj K Chaudhury. Roles of discontinuities in bio-inspired adhesive pads. Journal of the Royal Society Interface, 2(2):55–61, 2005.
  • [37] Shuman Xia, Laurent Ponson, Guruswami Ravichandran, and Kaushik Bhattacharya. Toughening and asymmetry in peeling of heterogeneous adhesives. Physical Review Letters, 108(19):196101, 2012.
  • [38] S.M. Xia, L. Ponson, G. Ravichandran, and K. Bhattacharya. Adhesion of heterogeneous thin films-I: Elastic heterogeneity. Journal of the Mechanics and Physics of Solids, 61(3):838–851, 2013.
  • [39] SM Xia, Laurent Ponson, Guruswami Ravichandran, and Kaushik Bhattacharya. Adhesion of heterogeneous thin films ii: Adhesive heterogeneity. Journal of the Mechanics and Physics of Solids, 83:88–103, 2015.
  • [40] Hong-Ping Zhao, Yecheng Wang, Bing-Wei Li, and Xi-Qiao Feng. Improvement of the peeling strength of thin films by a bioinspired hierarchical interface. International Journal of Applied Mechanics, 5(02):1350012, 2013.
  • [41] Animangsu Ghatak. Peeling off an adhesive layer with spatially varying topography and shear modulus. Physical Review E, 89(3):032407, 2014.
  • [42] Zhilong Peng and Shaohua Chen. Peeling behavior of a thin-film on a corrugated surface. International Journal of Solids and Structures, 60:60–65, 2015.
  • [43] D. Maugis. Contact adhesion and rupture of elastic solids. Solid State Sciences. Springer, 2000.